Skip to main content

Full text of "Julius Robert Oppenheimer Archive (PDF Library)"

See other formats


J. Robert 
Oppenheimer 
Science and the 
Common 
Understanding 


4h 


SCIENTIFIC— General Science 


The 1953 Reith Lectures, delivered over the B.B.C. from London, were 
directed by Professor Oppenheimer to an examination of “what there 
is new in atomic physics that is relevant, helpful, and inspiriting for 
men to know.” This led him, he says, into a relatively quiet room, that 
we know as quantum theory or atomic theory, in the house called 
“science.” 

“It is a vast house indeed. It does not appear to have been built upon 
any plan but to have grown as a great city grows. There is no central 
chamber, no one corridor from which all the others debouch. All about 
the periphery men are at work studying the vast reaches of space and 
the state of affairs billions of years ago; studying the intricate and 
subtle but wonderfully meet mechanisms by which life proliferates, 
alters and endures; studying the reach of the mind and its ways of 
learning; digging deep into the atoms and their unfathomed order. It is 
a house so vast that none of us know it, and even the most fortunate 
have seen most rooms only from the outside or by a fleeting passage, 
as in a king's palace open to visitors. It is a house so vast that there is 
not and need not be complete concurrence on where its chambers 
stop and those of the neighboring mansions begin. 

“It is not arranged in a line nor a square nor a circle nor a pyramid, 
but with a wonderful randomness suggestive of unending growth and 
improvisation. Not many people live in the house, relatively speaking— 
perhaps if we count all its chambers and take residence requirements 
quite lightly, one tenth of one per cent of all the people in this world— 
probably, by any reasonable definition, far fewer. And even those 
live elsewhere also, live in houses where the rooms are not lab = 
atomic theory or genetics or the internal constitution of the stars 


= 

quite different names like power and production and evil and Sy = 
and history and children and the word of God. —— 
“We go in and out; even the most assiduous of us is not bound tt€@ = 
vast structure. One thing we find throughout the house: there ar = 
locks; there are no shut doors; wherever we go there are the signs = 
usually the words of welcome. It is an open house, open to all comes = 
-- 


design: WARD & SAKS 


Science and the 


Common Understanding 


J. ROBERT OPPENHEIMER 


Simon and Schuster 


New York 


ALL RIGHTS RESERVED 
INCLUDING THE RIGHT OF REPRODUCTION 
IN WHOLE OR IN PART IN ANY FORM 
COPYRIGHT, 1953, 1954, BY J. ROBERT OPPENHEIMER 
PUBLISHED BY SIMON AND SCHUSTER, INC. 
ROCKEFELLER CENTER, 630 FIFTH AVENUE 
NEW YORK 20, N. Y. 


FIRST PAPERBACK PRINTING, 1966 


Library of Congress Catalog Card Number: 54-8650 
Dewey Decimal Classification Number: 530.1 


CONTENTS 


1 
NEWTON: THE PATH OF LIGHT 3 
2 
SCIENCE AS ACTION: RUTHERFORD’S WORLD 20 
3 
A SCIENCE IN CHANGE 35 
4 
ATOM AND VOID IN THE THIRD MILLENNIUM 51 
5 
UNCOMMON SENSE 68 
6 
THE SCIENCES AND MAN’S COMMUNITY 83 
APPENDIX I 101 
118 


APPENDIX II 


Science and the 


Common Understanding 


NEWTON: 


THE PATH OF LIGHT 


SCIENCE has changed the conditions of man’s life. 
It has changed its material conditions; by changing them 
it has altered our labor and our rest, our power, and the 
limits of that power, as men and as communities of men, 
the means and instruments as well as the substance of our 
learning, the terms and the form in which decisions of 
right and wrong come before us. It has altered the com- 
munities in which we live and cherish, learn and act. It has 
brought an acute and pervasive sense of change itself into 
our own life’s span. The ideas of science have changed the 
way men think of themselves and of the world. 

The description of these changes is not simple; it is rich 
in opportunity for error. As for the great material changes 
which science and practical art have made possible—ma- 
chines, for instance, or power, the preservation of life, the 
urbanization of populations, new instruments of war, new 
means of communication and information—these are but 


4 SCIENCE AND THE COMMON UNDERSTANDING 


part of the materials for the analysis of political economy 
and the wisdom and the insight of history. These are 
strands in the tangled affairs of men, and their evaluation 
is no more likely to be final and exhaustive than in any 
other part of history. 

As for the more direct effects of discovery in science on 
the way men think about things which are not themselves 
part of science, the historian of ideas has a similar prob- 
lem. Noting what in actual fact men have said about what 
they thought, who it was that thought it, and why he 
thought it, one finds, as in all history, that the contingent 
and the unpredictable, the peculiar greatnesses and blind- 
nesses of individual men play a determining part. One 
even finds the science of great scientists taken in the name 
of those scientists for views and attitudes wholly foreign 
and sometimes wholly repugnant to them. Both Einstein 
and Newton created syntheses and insight so compelling 
and so grand that they induced in professional philoso- 
phers a great stir of not always convenient readjustment. 
Yet the belief in physical progress, the bright gaiety, and 
the relative indifference to religion characteristic of the 
enlightenment, were as foreign to Newton’s character and 
preoccupation as could be; this did not keep the men of 
the enlightenment from regarding Newton as their patron 
and prophet. The philosophers and popularizers who have 
mistaken relativity for the doctrine of relativism have con- 
strued Einstein’s great works as reducing the objectivity, 
firmness, and consonance to law of the physical world, 
whereas it is clear that Einstein has seen in his theories of 


NEWTON: THE PATH OF LIGHT § 


relativity only a further confirmation of Spinoza’s view 
that it is man’s highest function to know and to understand 
the objective world and its laws. 

Often the very fact that the words of science are the 
same as those of our common life and tongue can be more 
misleading than enlightening, more frustrating to under- 
standing than recognizably technical jargon. For the 
words of science—relativity, if you will, or atom, or muta- 
tion, or action—have been given a refinement, a precision, 
and in the end a wholly altered meaning. 

Thus we may well be cautious if we inquire as to 
whether there are direct connections, and if so of what 
sort, between the truths that science uncovers and the way 
men think about things in general—their metaphysics— 
their ideas about what is real and what is primary; their 
epistemology—their understanding of what makes human 
knowledge; their ethics—their ways of thinking, talking, 
judging, and acting in human problems of right and 
wrong, of good and evil. 

These relations, the relations between scientific findings 
and man’s general views, are indeed deep, intimate, and 
subtle. If I did not believe that, I should hardly be ad- 
dressing these lectures to an attempt to elucidate what 
there is new in atomic physics that is relevant, helpful, 
and inspiring for men to know; but the relations are not, 
I think, relations of logical necessity. This is because sci- 
ence itself is, if not an unmetaphysical, at least a non- 
metaphysical activity. It takes common sense for granted 
as well as most of what has gone before in the specialized 


6 SCIENCE AND THE COMMON UNDERSTANDING 


sciences, And where it adds, alters, or upsets, it does so on 
the basis of an uncritical acceptance of a great deal else. 
Thus, to the irritation of many, the assertions of science 
tend to keep away from the use of words like “‘real”’ and 
“ultimate.” The special circumstances of the discovery of 
scientific truth are never very far from our minds when 
we expound it, and they act as a protecting sheath against 
their unlimited and universal acceptance. A few illustra- 
tions may make this clearer. 


We have discovered atoms. In many ways they act like 
the atoms of the atomists. They are the stuff of which mat- 
ter is made; their constellation and motion account for 
much—in fact, for most of the ordinarily observable prop- 
erties of matter. But neither they nor the smaller, less 
composite particles of which they are made are either per- 
manent, unchanging, or unchangeable. They do not act 
like objects of fixed form and infinite hardness. Such find- 
ings may be persuasive in discouraging the view that the 
world is made of fixed, immutable, infinitely hard little 
spheres and other shapes; but such findings are not in the 
nature of things conclusive, for one may always hold that 
the true atoms, the immutable, hard atoms, have so far 
eluded physical discovery, but that they are nevertheless 
there, and only when they are found will physics be deal- 
ing with the ultimate reality. Beyond that, one can hold 
that, although they may never be found by physical ex- 
periment, they are the underlying reality in terms of which 


NEWTON: THE PATH OF LIGHT 7 


all else, including the world of physics, is to be under- 
stood. 

Or, again, we may have discovered that as the nervous 
impulses pass from the retina of the eye toward the brain 
itself their geometric disposition resembles less and less 
that of the object seen. This may complicate or qualify the 
view that the idea is a geometric replica of the object of 
vision. It cannot and need not wholly exorcise it. 

The scientist may be aware that, whatever his findings, 
and indeed whatever his field of study, his search for truth 
is based on communication with other people, on agree- 
ment as to results of observation and experiment, and on 
talking in a common tongue about the instruments and ap- 
paratus and objects and procedures which he and others 
use. He may be aware of the fact that he has learned al- 
most everything he knows from the books and the deeds 
and talk of other people; and, in so far as these experi- 
ences are vivid to him and he is a thoughtful man, he may 
be hesitant to think that only his own consciousness is real 
and all else illusion. But that view, too, is not by logic 
exorcised; from time to time it may rule his spirit. 

Although any science gives countless examples of the 
interrelation of general law and changing phenomena, and 
although the progress of science has much to do with the 
enrichment of these relations, knowledge of science and 
practice of it and interest in it neither compel nor deny the 
belief that the changing phenomena of the actual world 
are illusion, that only the unchanging and permanent 
ideas are real. 


8 SCIENCE AND THE COMMON UNDERSTANDING 


If, in the atomic world, we have learned—as we have 
learned—that events are not causally determined by a 
strict, efficient, or formal cause; if we have learned to live 
with this and yet to recognize that for all of the common 
experience with ordinary bodies and ordinary happenings 
this atomic lack of causality is of no consequence and no 
moment, neither the one finding nor the other ensures that 
men when they think of the world at large are bound to a 
causal or a non-causal way of thinking. 

These many examples show that there can indeed be 
conflict between the findings of science and what a philoso- 
pher or a school of philosophy has said in great particular 
about some part of experience now accessible to science. 
But they also show that, if there are relationships between 
what the sciences reveal about the world and how men 
think about those parts of it either not yet or never to be 
explored by science, these are not relationships of logical 
necessity; they are not relationships which are absolute 
and compelling, and they are not of such a character that 
the unity and coherence of an intellectual community can 
be based wholly upon it. 

But if these examples indicate, as we should indeed ex- 
pect from the nature and conditions of scientific inquiry, 
that what science finds does not and cannot uniquely deter- 
mine what men think of as real and as important, they must 
show as well that there is a kind of relevance—a relevance 
which will appear different to different men and which 
will be responsive to many influences outside the work of 
science. This relevance is a kind of analogy, often of great 


NEWTON: THE PATH OF LIGHT 9 


depth and scope, in which views which have been created 
or substantiated in some scientific enterprise are similar 
to those which might be held with regard to metaphysical, 
epistemological, political, or ethical problems. The suc- 
cess of a critical and sceptical approach in science may en- 
courage a sceptical approach in politics or in ethics; the 
discovery of an immensely successful theory of great 
scope may encourage the quest for a simplified view of 
human institutions. The example of rapid progress in un- 
derstanding may lead men to conclude that the root of evil 
is ignorance and that ignorance can be ended. 

All these things have happened and all surely will hap- 
pen again. This means that, if we are to take heart from 
any beneficent influence that science may have for the 
common understanding, we need to do so both with mod- 
esty and with a full awareness that these relationships are 
not inevitably and inexorably for man’s good. 

It is my thesis that generally the new things we have 
learned in science, and specifically what we have learned 
in atomic physics, do provide us with valid and relevant 
and greatly needed analogies to human problems lying 
outside the present domain of science or its present bor- 
derlands. Before I talk of what is new I shall need to 
sketch, with perhaps an exaggerated simplicity and con- 
trast, the state of knowledge and belief to which these cor- 
rectives may apply. In doing this, we may have in mind 
that the general notions about human understanding and 
community which are illustrated by discoveries in atomic 
physics are not in the nature of things wholly unfamiliar, 


10 SCIENCE AND THE COMMON UNDERSTANDING 


wholly unheard of, or new. Even in our own culture they 
have a history, and in Buddhist and Hindu thought a more 
considerable and central place. What we shall find is an 
exemplification, an encouragement, and a refinement of 
old wisdom. We shall not need to debate whether, so 
altered, it is old or new. 

There are, then, two sketches that I would like to draw 
of the background for the altered experience of this cen- 
tury. One is the picture of the physical world that began 
to take shape in the years between Descartes’ birth and 
Newton’s death, that persisted through the eighteenth cen- 
tury, and with immense enrichments and extensions still 
was the basic picture at the beginning of our own. 

The second sketch has to do with the methods, the hopes, 
the program, and the style which seventeenth- and eight- 
eenth-century science induced in men of learning and in 
men of affairs, with some of the special traits of that pe- 
riod of enlightenment which we recognize today as so deep 
in our tradition, as both so necessary to us and so inade- 
quate. 

More than one great revolution had ended and had been 
almost forgotten as the seventeenth century drew its pic- 
ture of the physical world. A centuries-long struggle to 
decide whether it were rest or uniform motion that was the 
normal state of an undisturbed body no longer troubled 
men’s minds: the great clarity, so foreign to everyday ex- 
perience, that motion, as long as it was uniform, needed 
no cause and no explaining was Newton’s first law. The 
less deep but far more turbulent Copernican revolution 


NEWTON: THE PATH OF LIGHT 11 


was history: the earth revolved about the sun. The physi- 
cal world was matter in motion: the motion was to be un- 
derstood in terms of the impetus or momentum of the 
bodies which would change only for cause, and of the 
force that was acting upon it to cause that change. This 
force was immediate and proximate. It produced a tend- 
ency for the impetus to change, and every course could be 
analyzed in terms of the forces deviating bodies from their 
uniform motions. The physical world was a world of dif- 
ferential law, a world connecting forces and motions at 
one point and at one instant with those at an infinitely near 
point in space and point of time; so that the whole course 
of the physical world could be broken down into finer and 
finer instants, and in each the cause of change assigned by 
a knowledge of forces. 

Of these forces themselves the greatest in cosmic affairs 
—that which governed the planets in the heavens and the 
fall of projectiles on earth—had been found by Newton 
in the general law of gravity. Was this, too, something that 
spread from place to place, that was affected only instant 
by instant, point by point; or was it a property given as a 
whole, an interaction somehow ordained to exist between 
bodies remote from one another? Newton was never to 
answer this question; but he, and even more than he, Huy- 
gens, studying the propagation of light, were laying the 
foundations for a definite view—a view in which the void 
of the atomists would lose much of its emptiness and take 
on properties from the bodies which inhabited it, which in 
turn would affect bodies far away. 


12 SCIENCE AND THE COMMON UNDERSTANDING 


It was not until the nineteenth century and Faraday that 
the full richness of space began to be understood: how it 
could be the seat not only of gravitational forces produced 
by the mass of material particles but of electric and mag- 
netic forces produced by their charges. Even in Newton’s 
day it was clear that there were very strong forces at work 
in lending to material objects their solidity. Newton 
wrote: 


It seems probable to me, that God in the Beginning 
form’d Matter in solid, massy, hard, impenetrable, move- 
able Particles, of such Sizes and Figures, and with such 
other Properties and in such Proportion to Space, as most 
conduced to the End for which he formed them; and that 
these primitive Particles being Solids, are incomparably 
harder than any porous Bodies compounded of them; even 
so very hard, as never to wear or break in pieces; no ordi- 
nary Power being able to divide what God himself made 
one in the first Creation. 


Newton saw that what held atoms together and made mat- 
ter must be forces of inordinate strength, and he never con- 
sidered their existence without a sense of mystery and 
awe. He did not know, nor do we today know, in what 
subtle way these forces might or might not be related to 
the forces of gravity. 

But for many of his contemporaries and successors 
these questions appeared less pressing than the confidence 
that, once given the forces, the course of nature could be 


NEWTON: THE PATH OF LIGHT 13 


foretold and that, where the laws of gravity could be 
found, other forces would yield to observation and analy- 
sis. It is only in this century that we have begun to come 
to grips with other instances of antinomy, the apparent 
irreconcilability between the differential description of 
nature, point by point, instant to instant, and the total 
unique law and event. It is only in this century that we 
have had to recognize how unexpected and unfamiliar that 
relation between bodies and the atoms on the one hand, 
and that space full of light and electricity and gravita- 
tional forces on the other, could prove to be. 

For the eighteenth century the world was a giant 
mechanism. It was a causal world, whether or not gravity 
and the other forces acting on bodies inhered in them by 
their nature or by God’s will or that they, too, grew, 
through laws as rigorous as the laws of motion, from the 
properties induced in space by the bodies in it. All that 
happened had its full, complete, immediate, efficient 
cause. The great machine had a determinate course. A 
knowledge of its present and therefore its future for all 
time was, in principle, man’s to obtain, and perhaps in 
practice as well. These objects with which the world was 
filled—the heavenly bodies, the impenetrable atoms and 
all things composed of them—were found by observation 
and by experiment; but it would have occurred to no one 
that their existence and their properties could be qualified 
or affected by the observations that told of them. The giant 
machine was not only causal and determinate; it was ob- 


14 SCIENCE AND THE COMMON UNDERSTANDING 


jective in the sense that no human act or intervention 
qualified its behavior. 

A physical world so pictured could not but sharpen the 
great gulf between the object and the idea. It would do 
much to bring about that long, critical, and, in its later 
phase, irrational and mystical, view of the relations be- 
tween the knower and the known that started with Locke 
and is perhaps even today not fully or happily ended. 

It is, of course, clear that many developments in sci- 
ence that were to flower in the eighteenth and nineteenth 
centuries would soon moderate and complicate the harsh 
basic picture of the giant machine and of the vast gulf be- 
tween it and the knowing human mind that thought about 
it and analyzed its properties. This is true of the great de- 
velopment of statistics, which in the end made room for 
human ignorance as an explicit factor in estimating the be- 
havior of physical forces. It is true of chemistry, whose 
phenomena, whatever their ultimate description, looked so 
very little like the result of matter in motion. It is even 
more true of the biological sciences, where matter in mo- 
tion, ever evident and inevitable, appears both at first 
sight and upon deeper analysis only marginally relevant 
to what makes biological forms interesting. 

But with all of this, and with varying degrees of agree- 
ment and reservation, there was the belief that in the end 
all nature would be reduced to physics, to the giant ma- 
chine. Despite all the richness of what men have learned 
about the world of nature, of matter and of space, of 
change and of life, we carry with us today an image of the 


NEWTON: THE PATH OF LIGHT 15 


giant machine as a sign of what the objective world is 
really like. 

This view of the Newtonian world is oversimplified; 
perhaps any view of what men made of their new sciences, 
their new powers, and their new hopes will be simplified 
to the point of distortion. Science for the eighteenth cen- 
tury was not a finished undertaking; and, if men were 
overwhelmed with what they had learned, they were easily 
reminded of how much was still missing. A rational un- 
derstanding of the world was not an understanding for one 
generation or one man, as it is alleged that it at one time 
appeared to be to Descartes. The immense discoveries of 
the recent past made it impossible to hold the view that all 
that was really worth knowing had long been known—a 
view that is a sort of parody, in any case, of the Renais- 
sance. 

This was a long journey on which men were embarked, 
the journey of discovery; they would need their wits and 
their resources and their forbearance if they were to get 
on with it. But it was a job in which progress was inevita- 
ble, and in which the style and success of physical science 
would tend to set the style for all undertakings of man’s 
reason. What there is of direct borrowing from Newtonian 
physics for chemistry, psychology, or politics is mostly 
crude and sterile. What there is in eighteenth-century po- 
litical and economic theory that derives from Newtonian 
methodology is hard for even an earnest reader to find. 
The absence of experiment and the inapplicability of New- 
tonian methods of mathematical analysis make that inevi- 


16 SCIENCE AND THE COMMON UNDERSTANDING 


table. These were not what physical science meant to the 
enlightenment. 

It meant a style of thought, a habit of success, and an 
understanding of community quite typical for the age. 
These are to be found best in the learned communities that 
grew up in Europe and later in America—in the Royal 
Society and in the far more ambitious, far more revolu- 
tionary, far more programmatic French Academy. These 
communities were infused by a confidence in the power 
of reason and by a sense of improvement constant and al- 
most inevitable in the condition of man’s knowledge, and 
therefore of his actions and his life. They rest on a con- 
sensus of men, often seeing with their own eyes the crucial 
experiment that was to test or to confirm a theory; on the 
common experience of criticism and analysis; on the wide- 
spread use of mathematical methods with all the assurance 
of objectivity and precision that they give us. These were 
communities banded together for the promotion of knowl- 
edge—critical, rapacious to correct error, yet tolerant 
from knowing that error is an inevitable step in acquiring 
new knowledge. These were communities proud of their 
broad, non-sectarian, international membership, proud of 
their styleand their wit, and with a wonderful sense of new 
freedom. One may recapture some sense of these com- 
munities from the writings of the time. The first history of 
the Royal Society is not truly a history but an apology, 
written when the society was only a few years old, explain- 
ing it, defending it against its critics. Bishop Sprat has 
this to say: 


NEWTON: THE PATH OF LIGHT 17 


Their Purpose is, in short, to make faithful Records of 
all the Works of Nature, or Art, which can come within their 
Reach; that so the present Age, and Posterity, may be able 
to put a Mark on the Errors, which have been strengthened 
by long Prescription; to restore the Truths, that have lain 
neglected; to push on those, which are already known, to 
more various Uses; and to make the way more passable, to 
what remains unreveal’d. This is the Compass of their 
Design... 

They have tried to put it into a Condition of perpetual 
Increasing, by settling an inviolable Correspondence be- 
tween the Hand and the Brain. They have studied, to make 
it not only an Enterprise of one Season, or of some lucky 
Opportunity; but a Business of Time; a steady, a lasting, a 
popular, an uninterrupted Work . . . 

It is to be noted, that they have freely admitted Men of 
different Religions, Countries, and Professions of Life. 
This they were oblig’d to do, or else they would come far 
short of the Largeness of their own Declarations. For they 
openly profess, not to lay the Foundation of an English, 
Scotch, Irish, Popish, or Protestant Philosophy; but a 
Philosophy of Mankind. 


Reading this today, we can hardly escape a haunting 
sense of its timeliness and a certain nostalgia at how little 
the texture of our life conforms to these agreeable and 
noble ideals. We cannot perhaps wholly forget how much 
these communities owed to the long centuries of Christian 
life and Christian tradition; how much that they then took 
for granted in their inquiries and thoughts, in their whole 


18 SCIENCE AND THE COMMON UNDERSTANDING 


style, derived from a way of life and a history which they 
were about to change beyond all recognition; and how 
deeply this, their program, could alter the very men and 
the very minds to whom their program would in time be- 
come entrusted. 

These, however, were not reflections to darken much the 
eighteenth century or to cast real shadows on that great 
path of light, that renewed hope of men for a growing and 
growingly rational comprehension of their world and of 
themselves. At the very end of the century in another land 
largely nourished and fathered by the enlightenment, a 
gentleman and patriot wrote a letter. He wrote in answer 
to a young friend inquiring about his present course of 
study. He wrote in the last days of the Directorate, when 
the course of history was diverging in alarming and im- 
mense ways from that charted by the men of the French 
Academy. He wrote it about two years before he was to 
assume the Presidency of the United States, there for over 
a century to raise more firmly than ever before the stand- 
ard of man’s freedom, his progress, and his rational na- 
ture. 


I am among those who think well of the human char- 
acter generally. I consider man as formed for society, and 
endowed by nature with those dispositions which fit him 
for society. I believe also, with Condorcet, as mentioned in 
your letter, that his mind is perfectible to a degree of which 
we cannot as yet form any conception . . . science can 
never be retrograde; what is once acquired of real knowl- 
edge can never be lost. To preserve the freedom of the 


NEWTON: THE PATH OF LIGHT 19 


human mind then and freedom of the press, every spirit 
should be ready to devote himself to martyrdom; for as 
long as we may think as we will, and speak as we think, the 
condition of man will proceed in improvement. The gen- 
eration which is going off the stage has deserved well of 
mankind for the struggles it has made, and for having ar- 
rested that course of despotism which had overwhelmed 
the world for thousands and thousands of years. If there 
seems to be danger that the ground they have gained will 
be lost again, that danger comes from the generation your 
contemporary. But that the enthusiasm which characterises 
youth should lift its parracide hands against freedom and 
science would be such a monstrous phaenomenon as I can- 
not place among possible things in this age and country. 


The writer of the letter was Thomas Jefferson. 


SCIENCE AS ACTION: 


RUTHERFORD’S WORLD 


IT IS inherent in the very notion of culture and of 
tradition that there is a cumulative aspect to human life. 
The past underlies the present, qualifies and moderates it, 
in some ways limits it and in some ways enriches it. We 
understand Shakespeare better for having read Chaucer, 
and Milton for having read Shakespeare. We appreciate 
Trevelyan more for knowing Thucydides. We see Cézanne 
with better eyes for having looked also at Vermeer, and 
understand much more in Locke for knowing Aristotle, St. 
Matthew for knowing Job. But in actual fact we rather 
seldom bring a knowledge of the earlier to our first ac- 
quaintance with the later; and if it is true that Job throws 
light on Matthew, it is also true that Matthew throws light 
on Job. We can understand a great deal of what is written 
today, knowing little explicitly of what has been written 
in the past. We can and do know a great deal of what 
Shakespeare means and intends without any knowledge 


SCIENCE AS ACTION: RUTHERFORD’S WORLD 21 


of those earlier men who altered and educated his sensi- 
bility. 

The cumulative character of science is very different 
and very much more essential. It is one of the reasons for 
the great difficulty of understanding any science in which 
one has not largely become an expert—the science of 
which Hobbes wrote: “Of that nature, as none can under- 
stand it to be, but such as in good measure have attayned 
it.” 

There are at least two reasons for this: one has to do 
with the relation of later discoveries in science to earlier, 
and the other with the use that is made of earlier work in 
science as an instrument of progress. When we find out 
something new about the natural world this does not super- 
sede what we knew before; it transcends it, and the tran- 
scendence takes place because we are in a new domain of 
experience, often made accessible only by the full use of 
prior knowledge. The work of Huygens and Fresnel on the 
wave properties of light is as necessary today as it ever 
was, although we know that there are properties of light 
which are left out in their account and their experience, 
properties which, in the context of atomic happenings, are 
decisive. Newton’s law of gravitation and his equations of 
motion apply to and underlie immense realms of physical 
experience and are not made wrong by the fact that in 
other and still vaster spheres they must be replaced by the 
broader laws of Einstein. The chemical theory of valency 
has been explained, elucidated, and, to some small degree, 
extended by an understanding, in terms of the behavior of 


22 SCIENCE AND THE COMMON UNDERSTANDING 


electrons and nuclei, of what goes on in chemical bonding; 
but the chemical theory of valency is not superseded and 
will presumably be used as long as man’s interest in 
chemistry continues. The foundations of solid fact and the 
laws which describe it persevere through the whole course 
of science, to be refined and adapted to new contexts but 
never to be ignored or cast out. 

But this is only a part of the story. It is a recurring ex- 
perience of scientific progress that what was yesterday an 
object of study, of interest in its own right, becomes today 
something to be taken for granted, something understood 
and reliable, something known and familiar—a tool for 
further research and discovery. Sometimes the new instru- 
ment which is used to extend experience is a natural phe- 
nomenon, only barely qualified or controlled by the ex- 
perimenter. We are familiar with the use of calcite crys- 
tals to produce two separate beams of polarized light. We 
know that the cosmic rays are both an object of investiga- 
tion in themselves and a tool of hitherto unparalleled 
power for probing the properties and transmutations of 
primordia] matter here on earth and in the laboratory. 
Sometimes past knowledge is embodied not in a natural 
phenomenon but in an invention, or in elaborate pyramids 
of invention, a new technology. 

There are many well known and major examples of 
technological development during the last war which have 
added to the instruments of the investigator of the physical 
and biological world. We may recall two. Microwave 
radar—the generation, control, and detection of elec- 


SCIENCE AS ACTION: RUTHERFORD’S WORLD 23 


tromagnetic waves of relatively very short length—played 
a heroic part in the Battle of Britain. In the years since, 
it has provided powerful new means of investigating 
atomic, molecular, and even nuclear problems from which 
in actual fact subtle discoveries have been made about the 
laws of interaction of electrons and protons and neutrons. 

The nuclear reactor embodies in its technology very re- 
cently acquired understanding of the fission processes in 
uranium and of the behavior of neutrons in their collisions 
with atomic nuclei; it is now an important tool whose con- 
trolled and well-understood radiations are telling us about 
properties of matter hitherto barely accessible. Artificially 
radioactive substances made in great profusion by atomic 
reactors enable us to follow the course of individual atoms 
in chemical and biological changes. In biology especially 
they may be an addition to our instrumental facilities and 
techniques comparable in importance with the microscope 
itself. 


It is an oversimplification to say that technologies based 
upon recently discovered natural phenomena are taken as 
wholly for granted and as wholly known, but this is es- 
sentially the truth. They are added to the experimenter as 
a good tool is added to the artisan; as the pencil in the 
writer’s hand ceases to be an object in itself and becomes 
almost a part of the writer; or as a horse under a good 
horseman becomes for the time being not an animal to be 
cared for and thought about but a part of the entity “‘horse- 


24 SCIENCE AND THE COMMON UNDERSTANDING 


man.” Thus what has been learned and invented in science 
becomes an addition to the scientist, a new mode of per- 
ception, a new mode of his action. 

There are some cautions to be added to this. No experi- 
menter takes his equipment quite so much for granted that 
he fails to check whether in fact it is performing as it is 
supposed to perform; but the notion of how it is supposed 
to perform is for him in general a fixed thing not calling 
for further inquiry. This may be true even when the in- 
vention is a sample of practical art rather than a sample 
of true understanding. The photographic plate has served 
as an instrument of science for decades, during which its 
behavior was only very incompletely understood. Any 
machine can get out of order, and in a laboratory most ma- 
chines do. The horse is shod and bridled and fed before he 
can become part of the horseman. Nevertheless we use 
what we have learned to go further. A perpetual doubting 
and a perpetual questioning of the truth of what we have 
learned is not the temper of science. If Einstein was led to 
ask not ‘‘What is a clock?” but “How, over great dis- 
tances and with great precision, do we synchronize 
clocks?” that is not an illustration of the scepticism of 
science; it exemplifies rather the critical reason creating 
a new synthesis from paradoxes, anomalies, and bewilder- 
ments, which experiments carried on with new precision 
and in a new context brought into being. 

All this means that science is cumulative in a quite spe- 
cial sense. We cannot really know what a contemporary 
experiment means unless we understand what the instru- 


SCIENCE AS ACTION: RUTHERFORD’S WORLD 25 


ments and the knowledge are that are involved in its de- 
sign. This is one reason why the growing edge of science 
seems so inaccessible to common experience. Its findings 
are defined in terms of objects and laws and ideas which 
were the science of its predecessors. This is why the stu- 
dent spends many long years learning the facts and arts 
which, in the acts of science, he will use and take for 
granted—why this long tunnel, at the end of which is the 
light of discovery, is so discouraging for the layman to 
enter, be he an artist, scholar, or man of affairs. 

This conversion of an object of study into an instrument 
has its classic exemplification in Rutherford and the a- 
particle. This is a trail we will follow for some time. It 
will lead us to the heart of atomic physics. The a-particle, 
emitted by many naturally radioactive substances, identi- 
cal with the nucleus of helium, was indeed a strong right 
arm for Rutherford and all his school in probing the 
atomic world. Rutherford’s early works had been largely 
devoted to writing the wonderful natural history of the 
radioactive families—those which start with spontaneous 
changes in the heavy elements uranium and thorium. Part 
of the natural history was to discover the genetic relations 
between the various radioactive substances, some of them 
growing as a result of the decay of others and in turn giv- 
ing rise to daughter products by further transmutation. 

The natural history involves a chemical identification 
of the radioactive substances, the determination of the 
rapidity of their decay and of the alternative modes of 
decay, which some of them exhibit. It involves the recogni- 


26 SCIENCE AND THE COMMON UNDERSTANDING 


tion of three fundamentally different kinds of radiation, 
all of which appear at one stage or another in these family 
histories. This identification, which we shall meet again in 
later contexts, means learning some of the basic properties 
of the particles emitted. This identification, as we shall 
shortly see, is made possible by the fact that even a single 
such particle has readily detectable effects. 

These properties include the mass of the particle and its 
electric charge. These have usually been found in the first 
instance by studying their behavior in large-scale electric 
and magnetic fields and applying Newton’s laws to 
analyze their motion. These same methods give one a 
measure of the velocity or energy with which the particles 
are emitted, and of the loss of this energy as the particles 
pass through matter. Sometimes, at a later stage, the prod- 
ucts of an atomic or nuclear disintegration can be more 
thoroughly studied. They may have more subtle electro- 
magnetic properties than charge, such as a small magnetic 
moment. They may have structure or size. But the basic 
identifications can all be made in terms of the response 
of the radiations to familiar, large-scale, experimentally 
controllable situations like the classical electric and mag- 
netic fields of our laboratory courses. 

The a-particle of the naturally radioactive substances 
became for the middle years of Rutherford’s life the 
sharpest experimental tool; it was to be supplemented and 
to some extent superseded only when artificially accel- 
erated nuclei became available during the nineteen-thir- 
ties. The essential features of the experiments that have 


SCIENCE AS ACTION: RUTHERFORD’S WORLD 27 


told us most about atoms and nuclei and the ingredients of 
matter are two: one has to do with structure, and the other 
with scale. 

The structure of the experiment involves three parts: a 
probe, which is an object meant to explore or disturb mat- 
ter in its natural state, typically with some degree of vio- 
lence. This was the role of the a-particle. The second ele- 
ment is the target, which is some form of matter, whether 
pure or of controllable and manageable complications; 
and the third is the detector, which identifies and describes 
the objects emerging from the disturbance, whether they 
be the altered or the unaltered probe, or something 
knocked out of the target, or created in the collision, or 
something appearing long after the collision as evidence 
of a rearrangement of the collision products consequent 
upon the disturbance. This is not a universal pattern—this 
probe-target-detector assembly. The collision is not the 
only way of learning about atomic systems; but almost all 
of what we have learned has derived at least in part from 
such experiments and can be elucidated in terms of them. 

As to scale, it is the scale that determines the possibility 
of detection. The events that are so studied—the colli- 
sions, transmutations—can typically be studied event by 
event, atom by atom. The reason for this lies in two cir- 
cumstances: one is that in nuclear transformations, and 
even more so in transformations induced by cosmic rays 
and superaccelerators, the energy characteristic of a sin- 
gle atomic process is enormous compared to the chemical 
energies, and is sufficient to produce recognizable physical 


28 SCIENCE AND THE COMMON UNDERSTANDING 


and chemical changes in hundreds of thousands or even 
millions of atoms. 

The second circumstance lies in the art that has been de- 
voted toward exploiting these energies in systems of detec- 
tion. The detectors for Rutherford’s experiments are by 
now familiar. One is the scintillation screen, where an a- 
particle creates a flash of light easily visible through a 
microscope at the point where it hits the screen. Another is 
the beautiful cloud chamber of C. T. R. Wilson, which is, 
according to legend, an outcome of the inventor’s interest 
in the mist and clouds and rain of his native Scotland. In 
this cloud chamber the track of a charged particle is 
marked by the occurrence of innumerable small yet 
readily visible droplets of water or other liquid close to 
where the particles passed. A third is the counter, in which 
the electrical disturbance produced in a gas by the passage 
of a charged particle gives rise to a substantial electrical 
discharge, which can be amplified and analyzed by elec- 
tronic circuitry. 

These detectors have been supplemented by many, 
many others; and the precision and power of electronic 
amplification and analysis have been developed into a 
great art. The detector of atomic physics still characteristi- 
cally is designed to take advantage of the very great en- 
ergy involved in the changes of a single atom, and of the 
power to amplify this energy almost at will to make it ac- 
cessible. The clicking counters and flashing lights and oc- 
casionally even the ringing bells of a modern nuclear labo- 
ratory make the doings of individual atoms very vivid and 


SCIENCE AS ACTION: RUTHERFORD’S WORLD 29 


immediate, and make the subtle atoms of Epicurus or of 


Newton seem very private and remote. 


Rutherford and his probing a-particles and detectors 
are old history, dating back roughly some forty years. 
They are basic alike to atomic and nuclear physics, basic 
as a foundation for the great revolution in science which it 
is my principal purpose to describe, and for the further 
developments at the very forefront of contemporary dis- 
covery that have us:today perplexed and bemused. With 
his a-particles, obtained from natural radioactivity, Ruth- 
erford discovered the atomic nucleus and the nuclear 
model of the atom; with some help from other evidence he 
discovered the mass and the charge of the various atomic 
nuclei and thus rationalized Mendelyeev’s table of the ele- 
ments. With the a-particles, he was able to touch nuclear 
matter itself and measure its dimensions. He showed that 
it could be transmuted; he identified at least some of its 
ingredients. 

For the most part, a-particles when they pass through a 
bit of matter are not very much deflected or changed in di- 
rection; they are gradually slowed down; but occasionally 
a particle will change its direction of motion very greatly. 
It will be scattered through a large angle; it will act as 
though some great force had disturbed it, as though it had 
hit something quite small and quite hard. The law describ- 
ing these deflections is Rutherford’s law; and to it he gave 
a simple meaning: there are forces acting on the a-parti- 


30 SCIENCE AND THE COMMON UNDERSTANDING 


cles; they are not unfamiliar to physics. They are the elec- 
tric repulsion between the charge of the atomic nucleus 
and the charge of the a-particle—the same force which 
manifests itself when two positively charged pith balls 
push each other apart in an elementary demonstration. 
The balls repel each other because the two charges are 
similar; and the repulsion is described by Coulomb’s law 
—very much the same law as Newton’s law of gravitation. 
The repulsion is inversely proportional to the square of 
the separation of the charges. The charge of the atomic 
nucleus is a multiple of that of the proton—the nucleus of 
hydrogen. The multiple is the atomic number, which de- 
termines the number of electrons in the atom, and almost 
all the chemical properties of the element, and the position 
in the periodic table of that element. The mass of the 
nucleus is almost the whole mass of the atom as expressed 
by its atomic weight. This charge and mass is concentrated 
in a small volume. Everywhere outside it, the a-particle 
feels only the electric field. 

By using a-particles fast enough to overcome the elec- 
tric repulsion, and using light elements for which the 
charge and therefore the repulsion are not too great, Ruth- 
erford found that occasionally a-particles penetrated to a 
different domain entirely, where very strong forces, not 
electric forces, deflected them. In this way he found the 
dimensions of the nucleus itself: roughly one part in 10,- 
000 of the dimensions of the atom as a whole. This char- 
acterized the nucleus as a region of incredibly high den- 
sity, of many millions of tons per cubic inch. Rutherford 


SCIENCE AS ACTION: RUTHERFORD’S WORLD 31 


discovered even more: he was able to show that when fast 
a-particles penetrated nuclear matter things other than a- 
particles emerged from the mélée. In experiments under- 
taken during the First World War, and justified by Ruther- 
ford as of greater importance than any contribution he 
could then make to the prosecution of that war, Rutherford 
for the first time induced by human action the transmuta- 
tion of an atomic nucleus, knocking out of the nucleus of 
nitrogen a nucleus of hydrogen, or proton, and starting a 
chain of events which led, among many things, to man’s 
release of atomic energy, to what may some day be judged 
the most compelling argument of all for putting an end to 


war itself. 


The story went on from there. Before we revert to the 
nuclear model of the atom and how oddly different its 
properties are from any we can understand on the basis of 
Newtonian physics, we may follow sketchily and partially 
this course of discovery with probe, target, and detector 
that Rutherford initiated and that has continued until the 
present day. Twenty years ago, using the same a-particles 
as probes, Chadwick managed to identify another survivor 
of the disturbance, another ingredient of the nucleus, the 
neutron, which has roughly the proton’s mass but no 
charge, and thus to lay the foundations for an elementary 
view of nuclear composition. The nucleus is made up of 
neutrons and protons—enough protons to account for its 
charge, the atomic number; enough neutrons to account 


32 SCIENCE AND THE COMMON UNDERSTANDING 


for the excess of its atomic weight over its atomic number 
—held together in their tiny volume by strong forces 
wholly dissimilar from those of electricity and magnetism, 
whose description even today is a far from completely 
solved problem. 

Chadwick’s neutrons, in their turn, became probes, in- 
ducing nuclear transmutations very copiously, because 
they were not kept away from nuclei by the positive nu- 
clear charge. Their use led, in the years just before the 
war, to Hahn’s discovery that, when uranium was trans- 
muted by being hit by neutrons, among the products was 
barium, a large half of the original nucleus, but only 
about half—and thus to nuclear fission. 

Even this was only the beginning. In the very energetic 
particles of cosmic radiation, in the nuclei accelerated by 
giant modern accelerators to energies a hundredfold those 
of Rutherford’s a-particles, we have found new probes to 
elicit new phenomena; the story of sub-nuclear matter be- 
gan to unfold and ramify. A whole new family of hitherto 
unknown, and, for the most part, unrecognized and unex- 
pected objects began to emerge from the nuclear encoun- 
ters. The first of these were the various mesons, some 
charged and some uncharged, about ten times lighter than 
the proton and some hundreds of times heavier than the 
electron. In the last years there have appeared in increas- 
ing variety objects heavier than the mesons, other objects 
heavier even than protons, whose names are still being 
changed, from month to month, by solemn conferences. 
Physicists call them vaguely, and rather helplessly, “the 


SCIENCE AS ACTION: RUTHERFORD’S WORLD 33 


new particles.” They are without exception unstable, as in 
the neutron. They disintegrate after a time which varies 
from one millionth to less than a billionth of a second into 
other lighter components. Some of these components are 
in turn unfamiliar to physics and are themselves in turn 
unstable. We do not know how to give a clear meaning to 
this question. We do not know why they have the mass and 
charge that they do; why they and just they exist; why they 
disintegrate as they do; why in most cases they last as long 
as they do, or anything much about them. They are the 
greatest puzzle in today’s physics. ' 

But all this is now; and these were not the puzzles of 
Rutherford’s day. To these we shall turn in the next lec- 
tures. They become manifest when we try to deduce and 
describe the properties of Rutherford’s atom in terms of 
Newtonian mechanics. This attempted description failed. 
The atoms of nature are radically, dramatically, unlike 
atoms, composed as Rutherford found of electrons and 
small nuclei, subject to the forces Rutherford discovered 
and described, and moving according to Newton’s laws. 
The failure of this classical description turned out to be a 
major clue, one of the few major clues, in the atomic story. 
We learned, before the story was finished, that more than 
Newtonian mechanics would have to be modified if we 
were to understand and describe our experience with 
atomic systems. We would have to alter our ideas on very 
fundamental points, on causality, for instance, and even 
on the nature of the objectivity of parts of the physical 
world. We were to be reminded, in a quite unexpected 


34 SCIENCE AND THE COMMON UNDERSTANDING 


way, of the nature and limitations, as well as the power, 
of human knowledge itself. It is largely for this reason 
that the story of atomic discovery has appeared to me so 
full of instruction for us all, for layman as well as special- 
ist. For it has recalled to us traits of old wisdom that we 
can well take to heart in human affairs. 

Before these great changes could be completed, and the 
strange situation elucidated, many new ideas and methods 
of description were to be introduced. We learned words 
new for us, like “quantum,” and “‘state,”’ words like “‘cor- 
respondence” and “complementarity,” words with a new 
meaning for physics. Of these the word “‘correspondence” 
came to stand for the conservative and traditional traits of 
the new physics, that bound it to the physics of the past; 
whereas “complementarity” described, as we shall come 
to see, those new features, unknown to the physics of New- 
ton, that have broadened and humanized our whole under- 
standing of the natural world. 

Time and experience have clarified, refined, and en- 
riched our understanding of these notions. Physics has 
changed since then. It will change even more. But what we 
have learned so far, we have learned well. If it is radical 
and unfamiliar and a lesson that we are not likely to for- 
get, we think that the future will be only more radical and 
not less, only more strange and not more familiar, and that 
it will have its own new insights for the inquiring human 
spirit. 


A SCIENCE IN CHANGE 


OUR understanding of atomic physics, of what we 
call the quantum theory of atomic systems, had its origins 
at the turn of the century and its great synthesis and reso- 
lutions in the nineteen-twenties. It was a heroic time. It 
was not the doing of any one man; it involved the collabo- 
ration of scores of scientists from many different lands, 
though from first to last the deeply creative and subtle and 
critical spirit of Niels Bohr guided, restrained, deepened, 
and finally transmuted the enterprise. It was a period of 
patient work in the laboratory, of crucial experiments and 
daring action, of many false starts and many untenable 
conjectures. It was a time of earnest correspondence and 
hurried conferences, of debate, criticism, and brilliant 
mathematical improvisation. 

For those who participated, it was a time of creation; 
there was terror as well as exaltation in their new insight. 
It will probably not be recorded very completely as his- 


36 SCIENCE AND THE COMMON UNDERSTANDING 


tory. As history, its re-creation would call for an art as 
high as the story of Oedipus or the story of Cromwell, yet 
in a realm of action so remote from our common experi- 
ence that it is unlikely to be known to any poet or any his- 
torian. In other ways, there will be such times again. Most 
of us are convinced that today, in our present probings in 
the sub-atomic and sub-nuclear world, we are laying the 
groundwork for another such time for us and for our sons. 
The great growth of physics, the vast and increasingly com- 
plicated laboratories of the mid-twentieth century, the in- 
creasing sophistication of mathematical analysis, have al- 
tered many of the conditions of this new period of crisis. 
We do not think that they will have altered its heroic and 
creative character. 

When quantum theory was first taught in the universi- 
ties and institutes, it was taught by those who had partici- 
pated, or had been engaged spectators, in its discovery. 
Some of the excitement and wonder of the discoverer was 
in their teaching; now, after two or three decades, it is 
taught not by the creators but by those who have learned 
from others who have learned from those creators. It is 
taught not as history, not as a great adventure in human 
understanding, but as a piece of knowledge, as a set of 
techniques, as a scientific discipline to be used by the 
student in understanding and exploring new phenomena 
in the vast work of the advance of science, or its applica- 
tion to invention and to practical ends. It has become not 
a subject of curiosity and an object of study but an in- 
strument of the scientist to be taken for granted by him, 


A SCIENCE IN CHANGE 37 


to be used by him, to be taught to him as a mode of 
action, as we teach our children to spell and to add. 
What we must attempt to do in these talks is wholly dif- 
ferent. This is no school to learn the arts of atomic physics. 
Even those prior arts—the experimental tools, the mathe- 
matical powers, the theories, inventions, instruments, and 
techniques which defined the problems of atomic physics, 
which established the paradoxes, described the phenom- 
ena, and underlay the need for synthesis—are not known 
to us of our own experience. We must talk of our subject 
not as a community of specialized scientists but as men 
concerned with understanding, through analogy, descrip- 
tion, and an act of confidence and trust, what other people 
have done and thought and found. So men listen to ac- 
counts of soldiers returning from a campaign of unparal- 
leled hardship and heroism, or of explorers from the high 
Himalayas, or of tales of deep illness, or of a mystic’s 
communion with his God. Such stories tell little of what 
the teller has to tell. They are the threads which bind us in 
community and make us more than separate men. 


Here, then, we have our atoms. Their ingredients have 
been made manifest by Rutherford and his a-particles, as 
have the forces that act between the ingredients, and by 
probing with electrons and with light as well as with a-par- 
ticles. There is the nucleus, with almost the whole of the 
atom’s mass and almost none of its size, and with a charge 
which is measured by the atomic number, equal to the 


38 SCIENCE AND THE COMMON UNDERSTANDING 


number of electrons that surround the nucleus in the nor- 
mal atom. We have the simple laws of attraction and re- 
pulsion, familiar from the large-scale, everyday experi- 
ence with electricity. Unlike charges attract and like repel; 
and the forces, like Newton’s, decrease inversely with the 
square of their separation. 

In Rutherford’s day it seemed reasonable, as it no 
longer entirely does today in facing our modern physics, 
to subdivide the problem of atomic structure into three 
questions: what are the ingredients of the atom; what are 
the forces, and the laws of force, acting between these in- 
gredients; how in response to those forces do the ingredi- 
ents move? We know that even in atomic problems this 
division is not completely rigorous; but the refinements are 
minor and have largely proved tractable. They consist of 
taking into account the effect of the motion of the particles 
themselves on the forces between them and, in some cases, 
the distortion of the properties of particles, very small it- 
self in the atomic structures, by the presence of other par- 
ticles and the forces that they exert. It is surely not wholly 
true of the nucleus that these distortions are small; and in 
the strange objects which emerge so readily when nuclei 
undergo violent collisions we have persuasive, if indirect, 
evidence for that. 

The atom, then, has a massive charged nucleus; the 
atom as a whole is neutral and 10,000 to 100,000 times as 
far across as its tiny nuclear core. The rest of the atom is 
composed of electrons and electric fields—electrons that 
are the universal ingredients of matter, the determinants 


A SCIENCE IN CHANGE 39 


of almost all its chemical properties and of most of its 
familiar physical properties as well. There will be as many 
electrons in the atom as the atomic number, the nuclear 
charge; this makes the atom as a whole neutral. There will 
be one electron in hydrogen and thirteen in aluminum and 
ninety-two in uranium. These are the ingredients; and the 
laws of force, complex only in the last refinements, are 
basically simple. The electron feels an attractive Coulomb 
force exerted by the nucleus, attractive since the electron 
and nucleus are oppositely charged, and once again fall- 
ing off with distance in the same way as gravitational 
forces according to Newton’s law. For hydrogen, this 
means a simple situation: two bodies with a force between 
them identical in structure with that which the sun exerts 
on the planets; two bodies small enough compared to the 
atom’s size so that they almost never touch, and the prop- 
erties of their contact can have little influence. The law of 
forces has been verified not only by probing with particles, 
by which it was originally discovered, but by probing with 
electrons themselves, in the first instances by the beta rays 
of naturally radioactive substances. For other atoms there 
is in addition the electrical repulsion between the several 
electrons, balancing to some extent the nuclear attraction. 
And there is, further, the well-known mathematical com- 
plication of describing quantitatively the behavior of a 
system with many particles. 

But with hydrogen this should not be so. Here we have 
essentially a single light body moving in a simple and well- 
known force. The description of this system should be a 


40 SCIENCE AND THE COMMON UNDERSTANDING 


perfect example of Newtonian dynamics, and should, in 
its refinements, be intelligible in terms of all that the nine- 
teenth century had discovered about the behavior of 
charged particles in motion and the electromagnetic radia- 
tion produced when they are accelerated. 


But it did not turn out that way. To what appeared to 
be the simplest questions, we will tend to give either no 
answer or an answer which will at first sight be reminiscent 
more of a strange catechism than of the straightforward 
afirmatives of physical science. If we ask, for instance, 
whether the position of the electron remains the same, we 
must say “no”; if we ask whether the electron’s position 
changes with time, we must say “no”; if we ask whether 
the electron is at rest, we must say “no”; if we ask whether 
it is in motion, we must say “no.” The Buddha has given 
such answers when interrogated as to the conditions of a 
man’s self after his death; but they are not familiar an- 
swers for the tradition of seventeenth- and eighteenth-cen- 
tury science. 

Let us review, then, what a hydrogen atom should be 
like if we could apply Newton’s laws and the whole clas- 
sical picture of matter in motion to the simple model. The 
electron is held to its nucleus as the earth is to the sun, or 
as is Venus. It should revolve in an ellipse, as Kepler 
found and Newton explained. The size of the ellipse could 
be varied from atom to atom as the orbits of the planets 
are different, depending on how it was formed and what 


A SCIENCE IN CHANGE 41 


its history, and so should the shape of the orbits, whether 
they are narrow or round. There should be no fixed size 
for a hydrogen atom and no fixed properties; and when we 
disturb one by one of our probings, or when it is disturbed 
in nature, we would not expect it to return to a size and 
shape at all similar to that from which it started. This is 
not all—there are more recondite points. When a charge 
moves in anything but a straight line, it should send out 
electromagnetic radiation. This is what we see in every 
radio antenna. As far as our model goes, this radiation 
should in time sap the energy of the electron to make up 
for the energy that has been sent out in the form of light 
waves; and the ellipses on which the electron moves should 
get smaller and smaller as it gets nearer to its attractive 
sun and loses its energy. For a system about the size of the 
hydrogen atom as we know it in nature, a few hundredths 
of a millionth of an inch across, this process should go 
very rapidly; and the atom should become far, far smaller 
than atomic dimensions in very much less than a millionth 
of a second. The color of the light that the electron radiates 
should be determined by the period of its revolution; it too 
should be random, differing from orbit to orbit, differing 
from time to time as the orbits shrink and alter. This is the 
picture which classical physics—Newtonian physics—pre- 
dicts for the hydrogen atom, if Rutherford’s model is right. 

It could hardly be further from the truth. By all we 
know, hydrogen atoms if undisturbed are all identical. 
They are the same size and each has the same properties 
as any other, whatever its history, provided only that it has 


42 SCIENCE AND THE COMMON UNDERSTANDING 


had achance to recover from any disturbance. They last in- 
definitely. We think of them, rightly, as completely stable 
and unchanging. When they are undisturbed, they do 
not radiate light or any other electromagnetic radiation, as 
indeed they could not if they are to remain unaltered. 
When they are disturbed, they sometimes do radiate, but 
the color of the light that they emit is not random and con- 
tinuous but falls in the sharp lines of the hydrogen spec- 
trum. The very stability, extent, and definiteness is not at 
all understandable on the basis of classical physics; and 
indeed on the basis of classical physics there is no length 
that we can define in terms of the masses and charges of 
the ingredients of the atom, and that is even roughly of 
the actual dimensions of the atom. 

In other respects, too, the atomic system shows a pe- 
culiar lack of continuity wholly at variance with the prop- 
erties of Newtonian dynamics. If we probe atoms with a 
stream of electrons, for instance, the electrons will typi- 
cally lose some of their initial energy, but these losses are 
not random in amount. They correspond to definite, well- 
defined energy gaps, characteristic for the atom in ques- 
tion, reproducible and not too hard to measure. When 
an atom is irradiated by light, an electron will be ejected, 
if and only if the energy of that light exceeds a certain 
minimum knownas the photo-electric threshold. Indeed, it 
was this discovery which led Einstein in the early years of 
the century to a finding about light almost equally revo- 
lutionary for our understanding of light and for our un- 
derstanding of atomic systems. This finding, to be more 
precise, is that as one alters the frequency of the light that 


A SCIENCE IN CHANGE 43 


shines on a body, the energy of the electrons ejected in- 
creases linearly with the frequency; linearly—that means 
proportionally. The constant of proportionality, which 
connects energy with frequency, is the new symbol of the 
atomic domain. It is called Planck’s constant, or the quan- 
tum of action, and it gives a measure of energy in terms 
of frequency. It is the heraldic symbol over the gateway 
to the new world; and it led Einstein to the bold, though 
at the time hardly comprehensible, conclusion that light, 
which we know as an electromagnetic disturbance of 
rapidly changing. electrical fields, which we know as a 
continuous phenomenon propagating from point to point 
and from time to time like a wave, is also and is neverthe- 
less corpuscular, consisting of packets of energy deter- 
mined by the frequency of the light and by Planck’s con- 
stant. When a material system absorbs light, it absorbs 
such a packet, or quantum, of energy, neither less nor 
more; and the discontinuous nature of the energy ex- 
changes between an atom and an electron is paralleled by 
the discontinuous nature of the energy exchanged when 
radiation is absorbed or emitted. 

We shall have to come back more than once to light as 
waves, and light as quanta; but how radical a problem of 
understanding this presents can be seen at once from all 
of classical optics, from the work of Huygens and its 
mathematical elaboration by Fresnel, and even more com- 
pletely from its electromagnetic interpretation by Max- 
well. We know that light waves interfere. We know, 
that is, that if there are two sources of light, the 
intensity of the light to be found at some other place 


44 SCIENCE AND THE COMMON UNDERSTANDING 


will not necessarily be just composed of the sum of 
that which comes from the two sources; it may be more 
and it may be less. We know from unnumbered attempts 
how to calculate, and how to calculate correctly, what the 
interference of the sources will turn out to be. If we have 
light impinging on a screen which is opaque and there are 
two holes in the screen, not too large and not too far apart 
in the terms of the wave length, the wavelets that come 
from one of the holes will be added to those that come 
from the other. Where two crests of these wavelets coin- 
cide, we shall have more light than the sum of the two. 
Where a crest and a trough coincide, we shall have less; 
and so we observe and understand and predict and are 
quite confident of these phenomena of interference. 

Try for a moment to describe this in the terms of the 
passage of particles of quanta. If one of those quanta 
which characterize both the emission of light at the source 
and its detection—let us say, by the eye or by the photo- 
graphic plate or photo-cell on the far side of the screen— 
if a quantum passes through one of the holes, how can the 
presence of the other hole through which it did not pass 
affect its destiny? How can there be any science or any 
prediction if the state of affairs remote from the trajec- 
tory of the quantum can determine its behavior? Just this 
question and our slow answer to it will start us on the un- 
ravelling of the physics of the atomic world. 


The first great step, taken long before the crisis of quan- 
tum theory, was to find a way of describing atomic behav- 


A SCIENCE IN CHANGE 45 


ior, not forgetting the mechanics and electrodynamics of 
the past, but knowing that one had here to do with some- 
thing new and different, and necessarily postponing the 
question of the connection of that which is new with the old 
laws. This is Bohr’s first theory. It has given us the symbol 
of the atomic world: the nucleus and a series of circles and 
ellipses represent in a pictorial way the states of the atom. 
We use it today, though we know in far more detail and 
far more completely what Bohr knew when he proposed it, 
that it could be at best a temporary and partial analogy. 
This was Bohr’s first postulate: that in every atom there 
were stationary states whose stability and uniqueness 
could not be understood in terms of classical dynamics. 
The lowest one, the one with the least energy, the ground 
state, is truly stable. Unless we disturb it, it will last un- 
altered. The others are called excited states, and they may 
be excited by collision or radiation or other disturbance. 
They, too, are stable in a sense incomprehensible in terms 
of Newton’s theory. Their stability is not absolute though. 
Just as these states could be reached by transition induced 
by collision or disturbance, so an atom may return to states 
of lower energy, whether by further collision or spon- 
taneously. In these spontaneous changes it gives out that 
radiation which is the analog of the radiation which in 
classical theory would make all motion unstable. In simple 
cases, the energy of these stationary states and some of the 
properties such as their shape are identical with or similar 
to the energy of some of the properties of Newtonian or- 
bits. But this stops being true when we go even from hy- 


46 SCIENCE AND THE COMMON UNDERSTANDING 


drogen to helium, with its two electrons. It is only partially 
true in hydrogen; and the rules which Bohr laid down for 
determining the character of the orbits that would corre- 
spond to stationary states, the so-called quantum condi- 
tions, were from the first recognized by him as incomplete 
and provisional. We know now that the states are in fact 
nothing like orbits at all; that the element of change with 
time, which is inherent in an orbit, is missing from these 
states; and that in fact the very notion of an orbit can be 
applied to the motion of matter only when the stationary 
state is not defined, and that a stationary state can exist 
only when there is no possibility of describing an orbit at 
all. 

That was the first rule. And what is the second? The 
second rule is that an atom can change only by passing 
from state to state; that its energy changes by the differ- 
ence in energy between the states; and that, when this ex- 
change of energy occurs in the absorption, emission, or 
scattering of light, the frequency of the light will be re- 
lated to the energy by the relation of Einstein and of 
Planck. The energy will be the frequency multiplied by 
the quantum of action; thus atomic spectra directly reveal 
energy differences between states, and by this the whole 
field of spectroscopy becomes evidence for the location 
and the properties of atomic states, and we begin to learn 
what properties of these states are like those of classical 
orbits and what are unlike. 

But what are we to think of the transitions themselves? 
Do they take place suddenly? Are they very quick mo- 


A SCIENCE IN CHANGE 47 


tions, executed in going from one orbit to another? Are 
they causally determined? Can we say, that is, when an 
atom will pass from one of its states to another as we dis- 
turb it; and can we find what it is that determines that 
time? To all these questions, the answer would turn out 
to be “no.” What we learned to ask was what determined 
not the moment of the transition but the probability of the 
transition. What we needed to understand was not the state 
of affairs during the transition but the impossibility of 
visualizing the transition—an even more radical impos- 
sibility than with the states themselves—in terms of the 
motion of matter. We learned to accept, as we later 
learned to understand, that the behavior of an atomic 
system is not predictable in detail; that of a large number 
of atomic systems with the same history, in, let us say, the 
same state, statistical prediction was possible as to how 
they would act if they were let alone and how they would 
respond to intervention; but that nowhere in our battery of 
experimental probings would we find one to say what one 
individual atom would in fact do. We saw in the very heart 
of the physical world an end of that complete causality 
which had seemed so inherent a feature of Newtonian 


physics. 


How could all this be and yet leave the largely familiar 
world intact as we knew it? Large bodies are, of course, 
made up of atoms. How could causality for bullets and 
machines and planets come out of acausal atomic behav- 


48 SCIENCE AND THE COMMON UNDERSTANDING 


ior? How could trajectories, orbits, velocities, accelera- 
tions, and positions re-emerge from this strange talk of 
states, transitions, and probabilities? For what was true 
yesterday would be true still, and new knowledge could 
not make old knowledge false. Is there a possible unity 
between the two worlds and what is its nature? 

This is the problem of correspondence. Whatever the 
laws which determine the behavior of light or of electrons in 
atoms or other parts of the atomic world, as we come closer 
and closer to the familiar ground of large-scale experience, 
these laws must conform more and more closely to those 
we know to be true. This is what we call the principle of 
correspondence. In its formulation the key is the quantum 
of action, whose finiteness characterizes the new features 
of atomic physics. And so the physicist says that, where 
actions are large compared to the quantum of action, the 
classical laws of Newton and Maxwell will hold. What this 
tends to mean in practice is that when mass and distances 
are big compared to those of the electron and the atom’s 
size, classical theory will be right. Where energies are 
large and times long compared to atomic energies and 
times, we shall not need to correct Newton. Where this is 
so, the statistical laws of atomic physics will lead to proba- 
bilities more and more like certitudes, and the acausal fea- 
tures of atomic theory will be of no moment, and in fact 
lost in the lack of precision with which questions about 
large events will naturally be put. 

In Bohr’s hands and those of the members of his school, 
this correspondence principle was to prove a powerful 


A SCIENCE IN CHANGE 49 


tool. It did not say what the laws of atomic physics were, 
but it said something about them. They must in this sense 
be harmonious with, and ultimately reducible to, those of 
large-scale physics. And when to this principle was added 
the growing conviction that the laws of atomic physics 
must deal not with the Newtonian position, velocity, and 
acceleration that characterized a particle but with the ob- 
servable features of atoms—the energies and properties 
of stationary states, the probabilities of transitions between 
these states—the groundwork was laid for the discovery 
of quantum mechanics. 

The principle of correspondence—this requirement that 
the new laws of atomic mechanics should merge with those 
of Newtonian mechanics for large bodies and events—thus 
had great value as an instrument of discovery. Beyond 
that, it illustrates the essential elements of the relation of 
new discovery and old knowledge in science; the old 
knowledge, as the very means for coming upon the new, 
must in its old realm be left intact; only when we have left 
that realm can it be transcended. 

A discovery in science, or a new theory, even when it 
appears most unitary and most all-embracing, deals with 
some immediate element of novelty or paradox within the 
framework of far vaster, unanalyzed, unarticulated re- 
serves of knowledge, experience, faith, and presupposi- 
tion. Our progress is narrow; it takes a vast world unchal- 
lenged and for granted. 

This is one reason why, however great the novelty or 
scope of new discovery, we neither can, nor need, rebuild 


50 SCIENCE AND THE COMMON UNDERSTANDING 


the house of the mind very rapidly. This is one reason why 
science, for all its revolutions, is conservative. This is why 
we will have to accept the fact that no one of us really will 
ever know very much. This is why we shall have to find 
comfort in the fact that, taken together, we know more and 


more. 


ATOM AND VOID IN THE 


THIRD MILLENNIUM 


IN EXPLORING the atomic world, we have trav- 
eled to a new country, strange for those who have lived in 
the familiar world of Newtonian physics, strange even to 
Newton’s own view of wonder and pre-vision. “God in the 
Beginning,” he wrote, “form’d Matter in solid, massy, 
hard, impenetrable, movable Particles. . . .” 

We have our atoms; we are trying to understand them. 
We have the simplest of the atoms, hydrogen, with a single 
proton for its nucleus and a single electron to make it up. 
But the ingredients do not follow Newton’s laws of motion. 
Atoms of hydrogen appear to be all alike; they have a 
fixed size; they are stable and not transitory; the light that 
they emit is not what an electron circling in ever smaller 
ellipses would radiate. They have a stability that does not 
derive from Newtonian mechanics. When they are dis- 
turbed by light or electrons or other matter, they take up 
energy in definite quanta characteristic for the atom. They 


52 SCIENCE AND THE COMMON UNDERSTANDING 


are described in terms of states—states that are not orbits, 
though they have some of the properties of some special 
orbits. The states are stable, or almost stable. Transition 
between them, occasioned by disturbance, or occurring 
spontaneously with the emission of light, occurs by chance. 
We do not know the cause of the individual transition but 
only, at best, their probable distribution in time; nor do 
we have, in terms of space and time and trajectory, any 
picture whatever of what these transitions may be. These 
acausal atoms compose the familiar world of large bodies, 
orbits, and Newton’s laws. The laws that describe atomic 
behavior, the stationary states and transitions, reduce by 
correspondence, when applied to large systems, to New- 
ton’s laws. 

The discovery of these laws by Heisenberg could itself 
have led to all we now know of quantum theory, but it was 
supplemented as a matter of history by new discoveries in 
related fields which make the task of understanding and 
exposition simpler and more direct. Yet even these are 
both unfamiliar and abstract; I fear that no exposition can 
be wholly without difficulty, 


Our problem has to do with the so-called duality of 
wave and particle. On the one hand we have light, de- 
scribed in detail as a continuous electromagnetic wave 
with electric and magnetic fields, changing with a fre- 
quency that determines the light’s color, and with an am- 
plitude that determines its intensity. The waves of light 


ATOM AND VOID IN THE THIRD MILLENNIUM 53 


differ from radio waves only in one respect: their wave 
length is much shorter. They differ obviously from the 
waves we see on water, which are the more or less regular 
displacement of matter. But when we talk here of waves, 
in this account of wave-particle duality, as we shall have 
to, it will mean something quite abstract, something com- 
mon to light, radio, and water waves. 

It will mean a state of affairs distributed in space and 
propagating with time, sometimes a harmonic like a pure 
note of sound and sometimes irregular like noise. It will 
mean that these disturbances in general add, so that two 
crests reinforce, and a crest and a trough tend to cancel. It 
will mean that the sum of two effects may not be greater 
than either, but smaller, as the phases of crest and trough 
indicate. It will mean that, if we leave more than one al- 
ternative for a particle or for light to go from one place to 
another, the chance of arriving may be greater than the 
sum of the chances or smaller than the sum of the chances, 
because of this interference of the waves that represent the 
alternatives. 

When we deal with light, we deal with such waves; but 
we also deal, as Einstein discovered, with something sharp, 
discrete, and discontinuous—the light quantum. Whenever 
light acts on matter, or is produced by it, we find packets 
of defined energy and impulse, related to their frequency 
and their wave number by the universal proportionality 
of the quantum of action. How were these quanta to be 
thought of? Were they guided by the waves? Were they 
the waves? Were the waves an illusion, after all? 


54 SCIENCE AND THE COMMON UNDERSTANDING 


This turned out to be a universal quandary. De Broglie 
suggested and later Davison found that there were waves 
associated with electrons. Specifically Davison’s experi- 
ment showed that electrons, too, when they are scattered by 
the regular disturbance of a natural crystal, exhibit the 
same signs of interference, the same unmistakable signa- 
ture of the super-position of waves as light and as X-rays; 
and later experiments showed that this is true of all the 
other particles as well—protons, neutrons, and the atoms 
themselves. It would be true of large objects also were it 
not that their wave length is small, because of smallness of 
Planck’s constant, and becomes completely insignificant 
compared with their dimensions and with any practical 
possibility of determining their location and outline. 

All the questions which puzzled men about the relations 
of Einstein’s quanta and Maxwell’s waves were thus to be 
equally sharp and equally troublesome for the wave and 
particle properties of matter. The resolution of these ques- 
tions is the heart of atomic theory. They were brought to 
the point of crisis by another great discovery—Schroedin- 
ger’s discovery of his wave equation. 

In its original, bold form this was the discovery of a 
simple law for the propagation of electron waves—a natu- 
ral generalization of the connection between wave number 
and impulse, between energy and frequency, a generaliza- 
tion nevertheless adequate to describe the gross features of 
atomic systems and most of the familiar properties of mat- 
ter. This equation had many sorts of solutions. Some were 


ATOM AND VOID IN THE THIRD MILLENNIUM 55 


stationary, unchanging in time, with a frequency and en- 
ergy that corresponded to the stationary states of atoms. 
This same equation had other solutions of a very different 
kind, representing the trajectory of an electron as it might 
be seen crossing the Wilson cloud chamber. It had still 
other solutions, compounded by addition of several sta- 
tionary states with their several proper frequencies. These 
were not stationary but varied in time with frequencies 
corresponding to the spectrum of atoms and molecules. 

But what were these waves? What did they describe? 
How were they related to the ways in which we observe 
and study atomic systems, to Rutherford’s probings, to the 
collisions and disturbances of atoms? Schroedinger under- 
stood that in some sense the world of classical physics 
would emerge from his equation, whenever the wave 
lengths were small enough; then the trajectories for bodies 
and planets would be like the geometric paths of light, the 
rays of optics. But what would the waves mean when this 
was not the case? 

It would have been no answer to this question to attempt 
to interpret the waves as an essentially mechanical disturb- 
ance in some underlying mechanical medium; for the 
questions which needed answering had to do with the prob- 
lems of stationary states, and the behavior of electrons, 
and not with a sub-stratum inaccessible to observation. Nor 
was such a path followed. The discouraging outcome of an 
analogous attempt with electromagnetic waves was conclu- 
sive. It did not seem reasonable, nor in fact has it ever 


56 SCIENCE AND THE COMMON UNDERSTANDING 


proved possible, at a time when the very foundations of 
classical mechanics were being altered, to reinterpret this 
revolution in classical mechanical terms. 

There was another false start. It was at one time sug- 
gested that the waves, as they spread and moved, in some 
sense represented the changing shape, extension, and flow 
of the electron itself; when the disturbance grew larger the 
electron grew larger; when the wave moved faster the elec- 
tron moved faster. But to this interpretation there was an 
insuperable obstacle. Whenever we looked for the position 
of the particle, looked not directly with the eye, but with 
the natural extension of looking with a microscope, we did 
not find it spread out; we never found part of it in the 
place where we were looking. Either it was there or it was 
not there—the whole or none of it. Whenever we tried to 
measure the velocity of an electron or its impulse, we 
never found that part of it was moving with one speed and 
part with another; there was always one electron, one ve- 
locity, one answer to an experimental inquiry. The spread- 
ing of the waves in space thus did not mean that the elec- 
tron itself spread; it meant that the probability or likeli- 
hood of our finding the electron, when we look for it, 
spread as the wave does. 

And thus it was that these waves were recognized as de- 
scribing a state of affairs, as summarizing information we 
had about the electron, as very much more abstract waves 
indeed than we had hitherto encountered in physics. Their 
interpretation was statistical as well as abstract: where a 
disturbance was large, there we were likely to find the elec- 


ATOM AND VOID IN THE THIRD MILLENNIUM 57 


tron if we looked for it; where it was small, unlikely. If 
the disturbance had ripples in which a certain wave length 
was prominent, a measurement of the momentum would 
be likely to give us a value corresponding to that wave 
length. This clearly is qualitative talk. Quantitative rules 
for assigning a wave function to describe the outcome of 
an observation—or of other certain forms of knowledge, 
such as that of an atom in its state of lowest energy— 
needed to be, could be, and were developed; and they are 
a part of quantum theory. Their exposition presupposes 
some mathematical talk and calls at least for a blackboard. 
Similarly, the simple rules which relate the magnitude or 
properties of a wave function to the expectations that it 
implies for one or another observation are a rigorous and 
necessary part of the theory. But with these bonds to tie 
the wave to our knowledge and to interpret it for our pre- 
diction, the basis of the new physics has been laid. 


It is a statistical physics, as indeed might have been ex- 
pected from the statistical features of atomic transitions. 
Its predictions are in the form of assertions of probability 
and only rarely and specially in the form of certitudes. 
With this in mind, let us look again at our problem of 
interference, and of the two holes. 

Let us think of an opaque screen with two holes in it. 
Let us think of light, if we will; or, better still, let us think 
of electrons of a given velocity and therefore a given wave 
length and direction. We can do two experiments with a 


58 SCIENCE AND THE COMMON UNDERSTANDING 


source of electrons. In one, each hole in turn will be open 
for a little, while the other is closed; in the other, both 
holes will be open together. If we register the electrons on 
the far side of the screen, for instance, with a photographic 
plate, we see that the two patterns are radically different. 
In the one case, we have a transmission through each of 
the holes separately, with the characteristic diffraction 
pattern for that wave length and for holes of that diameter. 
These patterns are just added to one another on the photo- 
graphic film. But if both holes are open at the same time, 
something else happens. The waves that come through one 
interfere with those that come through the other; spots that 
were blackened before are now untouched and new spots 
appear where the electrons do arrive. 

If we try to think of this in terms of following the elec- 
trons through one or the other of the holes, we cannot un- 
derstand how it can make any difference whether that hole 
through which the electron did not pass is open or shut; yet 
it does. If we argue that the effect can be traced to the 
interaction of electrons passing through the two holes, we 
can disprove this by noting that the pattern is not affected 
by reducing the number of electrons to the point where 
there almost never are two passing through the two holes 
at the same time. What we are observing is something char- 
acteristic of the behavior of single particles, not of the 
interaction of several. 

Weare thus led to say that in this experiment a knowl- 
edge of which hole the electron passed through is in princi- 


ple inaccessible to us, that it is just the possibility of its 


ATOM AND VOID IN THE THIRD MILLENNIUM 59 


passing through one or the other that leads to the charac- 
teristic new interference phenomena, the new light spots 
and the new dark spots on the photographic film. We con- 
clude that, if we should make provision for registering 
through which hole the electron went, such as looking for 
it or observing the small push that it gives to the screen as 
it passes through, we would destroy the interference ef- 
fects. We would then have the same result as if we had in 
fact opened and shut the holes successively. 

We see the connection between these conclusions and the 
description of the state of affairs by a wave field in quali- 
tative terms, rather closely paralleling the arguments that 
were made quantitative in the uncertainty principle of 
Heisenberg. 

For we note that, if we were sure that the electron passed 
through one of the holes, the wave field would have to be 
restricted to that region; and that, if this were true, it 
would have to be composed not of a single wave length, or 
approximately a single one, but of waves of enough differ- 
ent wave lengths so that they can reinforce each other at 
one hole and vanish at the other; and we know that such 
waves have lost the coherent quality necessary for inter- 
ference. A little more generally, the waves of a single wave 
length will correspond to an electron of a definite velocity 
or impulse, but in an ill-defined or undefined position; the 
waves that are localized to represent a definition of posi- 
tion will be broadly scattered in wave length and represent 
an undefined velocity or impulse. This complementary re- 
striction on the degree to which a wave field can represent 


60 SCIENCE AND THE COMMON UNDERSTANDING 


both a well-defined position and a well-defined impulse is 
universal; it is measured by the quantum of action. It 
holds not only for electrons but for the more complicated 
waves that describe complex systems, for atoms and nu- 
clei and more composite bits of matter and more elemen- 
tary ones. And the very fact that no wave field can give 
that complete definition of the position and velocity of an 
object which was taken for granted in classical physics is 
also a description of the limitation on the observations 
which in the real world we shall manage to make. It rep- 
resents the fact that, when we study a system, making an 
experiment or an observation on it, we may—and in gen- 
eral we will, if we have prior knowledge before the experi- 
ment—be losing in whole or in part that prior knowledge. 
The experiment itself—that is, the physical] interactions 
between the system and the equipment that we are using 
to study it—will not only alter what we previously knew, 
but will in general alter it in a way which cannot be fol- 
lowed without invalidating the measurement or observa- 
tion we have undertaken. 

To cite but one example: if in the problem of the two 
holes we try to detect which hole the electron has passed 
through by noticing the push that it gives to the screen at 
that point, we shall have to leave a part of the screen free 
to respond to the push; and by this we lose all certitude as 
to where that part of the screen was when that electron 
passed through it. Many complex and detailed studies 
have been made of how this limitation of knowledge oc- 
curs in an experiment; but since the principle of comple- 


ATOM AND VOID IN THE THIRD MILLENNIUM 61 


mentarity, and the general adequacy of a wave field to 
describe a state of affairs, underlies the description of 
both the object and the instrument of observation, these 
examples only illustrate and make vivid what must gener- 
ally be true: the universal limitation, in contrast to classi- 
cal physics, of the extent to which all aspects of a physical 
system can be defined for the same system in the same 
instance. 

In observing atomic systems, in observing a system 
where the finiteness of the quantum of action plays an 
important part, we have a wide range of choice in the kind 
of probe, the kind of experiment, the kind of experimental 
equipment we wish to make. To any of these, if it is a good 
experiment, there will be a meaningful answer which tells 
us what the state of affairs is. From this, and from the 
wave field which represents it, we can then make statistical 
predictions of what will happen in a subsequent experi- 
ment. The potentialities of measurement are varied. We 
can do one thing or another; there are no inherent limits 
on the choice of actions on the part of the observer. 

This is a very different view of reality from Newton’s 
giant machine. It is not causal; there is no complete causal 
determination of the future on the basis of available 
knowledge of the present. The application of the laws of 
quantum theory restricts, but does not in general define, 
the outcome of an experiment. This means that every ob- 
servation on a system reveals some new knowledge as to 
what its state is that did not exist before, and could not by 
analysis and mathematical computation have been ob- 


62 SCIENCE AND THE COMMON UNDERSTANDING 


tained. It means that every intervention to make a meas- 
urement, to study what is going on in the atomic world, 
creates, despite all the universal order of this world, a 
new, a unique, not fully predictable, situation. 


Even in a brief account other points need to be men- 
tioned. We have almost lost the concept of equations of 
motion, having discovered that the very terms in which 
they are formulated—position, velocity, acceleration, and 
force—are not simultaneously applicable and do not, 
taken together, correspond to things that we know about 
the electron with enough accuracy to be meaningful for an 
atomic system. Instead, what we can have is a knowledge 
of the state, summarizing for us what we have found by 
observation; and the analogue of the equation of motion 
must tell us how, in response to forces acting within the 
system or upon it, this state will change with time. This, it 
turns out, is just what Schroedinger’s equation does. And 
once again this equation, when applied to the familiar con- 
texts of massive bodies and great distances, where the 
quantum of action is in fact negligibly small, will describe 
for us waves so reasonably concentrated in space, so little 
dispersed about their average wave length, that the New- 
tonian orbit reappears in its unaltered, classical path. 

But this condition—this emergence of an orbit—is a 
long way from the wave that describes the normal state 
of an atom. State and orbit, like position and impulse, are 
complementary notions; where one applies, the other can- 


ATOM AND VOID IN THE THIRD MILLENNIUM 63 


not be defined, and for a full description we must be able 
to use now one, now the other, depending on the observa- 
tion and the questions that we put. 

When we speak here of observer and object, of instru- 
ment or probe, and system to be probed, we are not talk- 
ing of the mind of man. We are talking of a division 
between the object of study and the means used to study 
it. That division can be made in more than one way. We 
may regard the a-particles that Rutherford used as an in- 
strument, and their response as a measure of the state of 
affairs. We may regard the a-particle as a part of the sys- 
tem we are studying, and the slits that define its path or 
the fields that deflect it and the screens that detect it as the 
instrument. But whichever we do, the observation will al- 
ways be transformed into some large-scale happening— 
some flash of light, some triggering of a circuit, some 
pointing of a pointer on the dial of an instrument—which 
is well defined and familiar and unambiguous, and where 
the question of our freedom to do one or another observa- 
tion on it no longer is relevant. The atomic world has not 
lost its objective quality; but it attains this by means of 
those interlockings with experiment which we use to define 
one or another of its properties and to measure them. 

It needs to be clear that what is described here is 
not an expression of mood or preference or taste; it is 
an exact, beautiful, quantitative, immensely versatile, 
and immensely successful science. It is what students 
learn when they prepare themselves for further re- 
searches in physics, or what engineers learn whose 


64 SCIENCE AND THE COMMON UNDERSTANDING 


engineering involves a knowledge of the solid state of 
physical materials, or what chemists learn if they wish to 
understand the subtler features of chemical bonding or 
chemical kinetics, or astronomers if they wish to know 
what things are like in the interior of the stars. One could 
go much farther in describing this discipline, even without 
mathematics; but the words would before long become 
cumbersome and unfamiliar and almost a misinterpreta- 
tion of what in mathematical terms can be said with beauty 
and simplicity. 


Even some of the more paradoxical features of quantum 
theory turn out to be related to practical matters of real 
importance. One of the earliest to be noted and the oddest 
is this: if, in familiar life, we roll a ball up a hill and it 
does not have enough vigor to get over the top, it will roll 
back on the same side; it will not get through the hill. But 
if we bombard sucha hill with a-particles or electrons they 
may have a small chance of getting through, even when 
they cannot get over. This has a close analogy with the fact 
that very small objects do not cast sharp shadows in a 
beam of light. Light because of its wave nature bends 
around them. It corresponds to the fact that when we let 
electrons or other particles of definite energy encounter 
a barrier, neither the kinetic nor the potential energy alone 
can be completely well defined; and indeed, were we to 
try to detect the electron just as it passes through the hill, 
we should need an experiment that could give the electron 


ATOM AND VOID IN THE THIRD MILLENNIUM 65 


enough energy to be quite legitimately on top of the hill. 
This penetration of barriers is not without importance. It 
accounts for the fact that the a-particles that Rutherford 
used could sometimes, after millions of years, escape from 
the nuclei through a high hill where electrostatic repulsion 
had imprisoned them. It accounts for the fact that in the 
sun and other stars nuclei having only very moderate en- 
ergy occasionally come into contact and react. Thus the 
stars light the heavens, and the sun warms and nourishes 
the earth. 

Another consequence of the wave-like character of all 
matter is that, when particles with very low velocity and 
very long wave length bombard other particles of matter, 
they may interact far more often than if these interactions 
were limited to their coming in contact. The very lack of 
definition of their relative position makes interaction pos- 
sible, in some cases over distances characterized not by 
their dimensions but by their wave length. This is the cir- 
cumstance which, among many others, enables the rare 
Uranium-235, as it occurs in natural uranium, to catch up 
enough of the neutrons which fly about to sustain a chain 
reaction in an atomic reactor. 

There are even some odd things about the identity and 
the identifiability of the electrons themselves. That they 
are all similar we know. Their inherent properties, their 
charge, their mass when at rest, are the same. We wish that 
we understood this better; some day, no doubt, we shall; 
but we know that it is true. But if classical physics were 
the whole story, we could still, if we wished, always iden- 


66 SCIENCE AND THE COMMON UNDERSTANDING 


tify an electron, and know that it was the same as the one 
we had seen before. We could follow it, not, it is true, 
without trouble, but without paradox, without inconsist- 
ency, from where we first found it through its collisions 
and interactions and deflections and changes by keeping 
in touch with its trajectory. If it hit another electron, we 
would know which it was that came out in one direction 
and which in another. In fact this is not really true, except 
in those special instances where the collision is of such low 
energy that the two electrons can be described by waves 
which never overlap at the same place at the same time. As 
soon as that is no longer the case, we lose in principle all 
ability to tell one electron from another; and in atomic 
physics, where the electrons of an atom, and even the elec- 
trons of neighboring atoms, are not well defined in position 
and can often occupy the same volume, we have no way 
of identifying the individual particle. This, too, has con- 
sequences. When two electrons collide, the wave that rep- 
resents one of them and the wave that represents the other 
may, and do, interfere; and this gives rise to novel effects 
and new forms for the interactions produced by their elec- 
tric repulsion. It is responsible for the permanent magnet- 
ism of magnets. It is responsible for the bonding of or- 
ganic chemistry and for the very existence in any form 
that we can readily imagine of living matter and of life 
itself. 

These examples are not given to perplex and bemuse. 
They are rather illustrations of how even the most para- 
doxical and unexpected consequences of the new mechan- 


ATOM AND VOID IN THE THIRD MILLENNIUM 67 


ics, of wave-particle duality, and of complementarity are 
involved in an understanding of important and familiar 
features of the natural world, and of how massive is the 
system of understanding and knowledge of which they are 
a part. 


UNCOMMON SENSE 


A CENTURY after Newton, in 1784, the progress 
of that century was celebrated in an anonymous memorial 
lodged in the ball of the tower of St. Margaret’s church at 
Gotha, to be found by men of future times. It read: 


“Our days have been the happiest time of the eight- 
eenth century. . . . Hatred born of dogma and the com- 
pulsion of conscience sink away; love of man and freedom 
of thought gain the upper hand. The arts and sciences blos- 
som, and our vision into the workshop of nature goes deep. 
Artisans approach artists in perfection; useful skills flower 
at all levels. Here you have a faithful portrait of our time. 
. . » Do the same for those who come after you and re- 
joice!” 


Transience is the backdrop for the play of human prog- 
ress, for the improvement of man, the growth of his know]- 
edge, the increase of his power, his corruption and his 


UNCOMMON SENSE 69 


partial redemption. Our civilizations perish; the carved 
stone, the written word, the heroic act fade into a memory 
of memory and in the end are gone. The day will come 
when our race is gone; this house, this earth in which we 
live will one day be unfit for human habitation, as the sun 
ages and alters. 

Yet no man, be he agnostic or Buddhist or Christian, 
thinks wholly in these terms. His acts, his thoughts, what 
he sees of the world around him—the falling of a leaf or a 
child’s joke or the rise of the moon—are part of history; 
but they are not only part of history; they are a part of 
becoming and of process but not only that: they partake 
also of the world outside of time; they partake of the light 
of eternity. 

These two ways of thinking, the way of time and history 
and the way of eternity and of timelessness, are both part 
of man’s effort to comprehend the world in which he lives. 
Neither is comprehended in the other nor reducible to it. 
They are, as we have learned to say in physics, comple- 
mentary views, each supplementing the other, neither tell- 
ing the whole story. Let us return to this. 


First, we had best review and extend somewhat this 
account of the complementarity of the physicists. In its 
simplest form it is that an electron must sometimes be 
considered as a wave, and sometimes as a particle—a 
wave, that is, with the continuous propagation and char- 
acteristic interference that we learn to understand in the 


7O SCIENCE AND THE COMMON UNDERSTANDING 


optics laboratory, or as a particle, a thing with well-de- 
fined location at any time, discrete and individual and 
atomic. There is this same duality for all matter and for 
light. In a little subtler form this complementarity means 
that there are situations in which the position of an atomic 
object can be measured and defined and thought about 
without contradiction; and other situations in which this 
is not so, but in which other qualities, such as the energy 
or the impulse of the system, are defined and meaningful. 
The more nearly appropriate the first way of thinking is to 
a situation, the more wholly inappropriate the second, so 
that there are in fact no atomic situations in which both 
impulse and position will be defined well enough to permit 
the sort of prediction with which Newtonian mechanics has 
familiarized us. 

It is not only that when we have made an observation on 
a system and determined, let us say, its position, we do not 
know its impulse. That is true, but more than that is true. 
We could say that we know the position of that system and 
that it may have any one of a number of different im- 
pulses. If we try on that basis to predict its behavior as a 
sort of average behavior of all objects which have the 
measured position and which have different and unmeas- 
ured impulses, and work out the average answer according 
to Newton’s laws, we get a result that is wholly at variance 
with what we find in nature. This is because of the peculiar 
property, which has no analogue in the mechanics of large 
objects, of interference between waves representing the 
consequences of assuming one impulse and those of assum- 


UNCOMMON SENSE 71 


ing another. We are not, that is, allowed to suppose that 
position and velocity are attributes of an atomic system, 
some of which we know and others of which we might know 
but do not. We have to recognize that the attempt to dis- 
cover these unknown attributes would lose for us the 
known; that we have a choice, a disjunction; and that this 
corresponds to the different ways we can go about observ- 
ing our atom or experimenting with it. 

We have a state of affairs completely defined by the 
nature of the observation and by its outcome—the nature 
determining what properties of the system will be well de- 
fined in the state and what poorly. The outcome then is the 
determination of the well-defined quantities by measure- 
ment. This state thus is a summary, symbolic and uncom- 
fortably abstract for general exposition, of what sort of 
observation we have made and what we have found 
through it. It codifies those characteristics of the experi- 
mental arrangement which are reliable, in the sense that 
the equipment we use records something that we know 
about atomic systems. It describes also those character- 
istics that are indeterminate, in the sense that they may not 
only have been disturbed or altered, but that their disturb- 
ance cannot be registered or controlled without the loss, in 
the experiment, of all ability to measure what was sup- 
posed to be measured. 

This state, this description of the atom, is not the only 
way of talking about it. It is the only way appropriate to 
the information we have and the means that we have used 
to obtain it. It is the full account of this information; and 


72 SCIENCE AND THE COMMON UNDERSTANDING 


if the experiment was properly and scrupulously done it 
tells us all that we can find out. It is not all that we could 
have found out had we chosen a different experiment. It is 
all that we could find out having chosen this. 

This state is objective. We can calculate its properties, 
reproduce it with similar atoms on another occasion, verify 
its properties and its ways of change with time. There is no 
element of the arbitrary or subjective. Once we have done 
our experiment and its result is recorded and the atom 
disengaged, we know its meaning and its outcome; we can 
then forget the details of how we got our information. 

But, although the state of the system is objective, a 
mechanical picture of how it was brought into being is not 
generally possible. There is a most vivid example of this, 
made famous by the prominent part it played in the de- 
bates between Einstein and Bohr as to the meaning and 
adequacy of atomic theory. It can be put rather simply. 
Let us suppose that we have two objects; one of them may 
be an electron or an atom, and it will be the one we wish 
to study. The other may be a relatively large piece of 
matter—a screen with a hole through it, or any other 
body; but it should be heavy so that its motion will be 
unimportant compared to that of the electron. Let us sup- 
pose that we by measurement know the impulse or mo- 
mentum of both of these objects, and have them collide. 
Let the electron go through the hole, or bounce off the 
other body. If, after the collision, we measure the impulse 
of the heavy body, we will then know that of the electron 
because, as Newton’s third law teaches us, the sum of the 


UNCOMMON SENSE 73 


impulses is not altered by the collision. In that case we 
would have a state of the electron of well-defined impulse, 
as precisely defined as we had made the precision of our 
measurements. If, on the other hand, we observed the posi- 
tion of the heavy body, we would know where the light one 
had been at the moment of the collision, and so would have 
a quite different description of its state, one in which its 
position and not its impulse had been well defined—or, in 
the language of waves, a spherical wave with its center at 
the point of collision, and not a plane wave with its direc- 
tion and wave length corresponding to the momentum. 

We have thus the option of realizing one or the other of 
two wholly dissimilar states for the electron, by a choice of 
what we observe about the heavy body with which it once 
was in interaction. We are not, in any meaningful sense, 
physically altering or qualifying the electron; we are de- 
fining a part of, although in this case a late part of, the 
experimental procedure, the very nature of the experiment 
itself. If we exercise neither option, if we let the heavy 
body go with unmeasured momentum and undefined posi- 
tion, then we know nothing of the electron at all. It has no 
state, and we are not prepared to make any meaningful 
predictions of what will become of it or of what we shall 
find should we again attempt an experiment upon it. The 
electron cannot be objectified in a manner independent of 
the means chosen for observing or studying it. The only 
property we can ascribe to it without such consideration 
is our total ignorance. 

This is a sharp reminder that ways of thinking about 


74 SCIENCE AND THE COMMON UNDERSTANDING 


things, which seem natural and inevitable and almost ap- 
pear not to rest on experience so much as on the inherent 
qualities of thought and nature, do in fact rest on experi- 
ence; and that there are parts of experience rendered ac- 
cessible by exploration and experimental refinement where 
these ways of thought no longer apply. 

It is important to remember that, if a very much subtler 
view of the properties of an electron in an atomic system 
is necessary to describe the wealth of experience we have 
had with such systems, it all rests on accepting without 
revision the traditional accounts of the behavior of large- 
scale objects. The measurements that we have talked about 
in such highly abstract form do in fact come down in the 
end to looking at the position of a pointer, or the reading 
of time on a watch, or measuring out where on a photo- 
graphic plate or a phosphorescent screen a flash of light or 
a patch of darkness occurs. They all rest on reducing the 
experience with atomic systems to experiment and observa- 
tion made manifest, unambiguous, and objective in the be- 
havior of large objects, where the precautions and incerti- 
tudes of the atomic domain no longer directly apply. So it 
is that ever-increasing refinements and critical revisions in 
the way we talk about remote or small or inaccessible parts 
of the physical world have no direct relevance to the fa- 
miliar physical world of common experience. 


Common sense is not wrong in the view that it is mean- 
ingful, appropriate, and necessary to talk about the large 


UNCOMMON SENSE 75 


objects of our daily experience as though they had a veloc- 
ity that we knew, and a place that we knew, and all the rest 
of it. Common sense is wrong only if it insists that what is 
familiar must reappear in what is unfamiliar. It is wrong 
only if it leads us to expect that every country that we 
visit is like the last country we saw. Common sense, as the 
common heritage from the millennia of common life, may 
lead us into error if we wholly forget the circumstances to 
which that common life has been restricted. 

Misunderstanding of these relations has led men to wish 
to draw from new discoveries, and particularly those in the 
atomic domain, far-reaching consequences for the ordi- 
nary affairs of men. Thus it was noted that, since the ulti- 
mate laws of atomic behavior are not strictly causal, not 
strictly determinate, the famous argument of Laplace for 
a wholly determinate universe could not be maintained. 
And there were men who believed that they had discovered 
in the acausal and indeterminate character of atomic 
events the physical basis for that sense of freedom which 
characterizes man’s behavior in the face of decision and 
of responsibility. 

In a similar light-hearted way it was pointed out that, 
as the state of an atomic system requires observation for 
its definition, so the course of psychological phenomena 
might be irretrievably altered by the very effort to probe 
them—as a man’s thoughts are altered by the fact that he 
has formulated and spoken them. It is, of course, not the 
fact that observation may change the state of an atomic 
system that gives rise to the need for a complementary 


76 SCIENCE AND THE COMMON UNDERSTANDING 


description; it is the fact that, if the observation is to be 
meaningful, it will preclude any analysis or control of that 
change, that is decisive. 

But these misapplications of the findings of atomic phys- 
ics to human affairs do not establish that there are no valid 
analogies. These analogies will, in the nature of things, be 
less sharp, less compelling, less ingenious. They will rest 
upon the fact that complementary modes of thought and 
complementary descriptions of reality are an old, long- 
enduring part of our tradition. All that the experience of 
atomic physics can do in these affairs is to give us a re- 
minder, and a certain reassurance, that these ways of talk- 
ing and thinking can be factual, appropriate, precise, and 
free of obscurantism. 

There are a number of examples which are illuminated 
by, and in turn illuminate, the complementarity of atomic 
theory. Some of them are from quite different parts of 
human life and some of them from older parts of science. 
There is one from physics itself which is revealing, both 
in its analogies and its points of difference. One of the 
great triumphs of nineteenth-century physics was the ki- 
netic theory of heat—what is called statistical mechanics. 
This is both an interpretation and a deduction of many of 
the large-scale properties and tendencies of matter: of the 
tendency, for instance, of bodies that can exchange heat to 
come to a common temperature, or of the density of a gas 
to be uniform throughout a container, or of work to dis- 
sipate itself in heat, or quite generally of all of those ir- 
reversible processes in nature wherein the entropy of 


UNCOMMON SENSE 77 


systems increases, and forms become more uniform and 
less differentiated when left to themselves to develop. 
The phenomena we deal with here are defined in terms 
of temperature and density and pressure and other large- 
scale properties. The kinetic theory, statistical mechanics, 
interprets the behavior of these systems in terms of the 
forces acting on the molecules and of the motion of the 
molecules that compose them, which are usually quite ac- 
curately described by Newton’s laws. But it is a statistical 
theory of this motion, recognizing that in fact we do not 
in general know, and are not in detail concerned with, the 
positions and velocities of the molecules themselves, but 
only with their average behavior. We interpret the temper- 
ature of a gas, for instance, in terms of the average kinetic 
energy of its molecules, and the pressure as the average of 
the forces exerted by the collision of these molecules on 
the surface of the container. This description in terms of 
averages, embodying as part of itself our ignorance of the 
detailed state of affairs, is thus in some sense complemen- 
tary to a complete dynamic description in terms of the 
motion of the individual molecules. In this sense kinetic 
theory and dynamics are complementary. One applies to a 
situation in which the individual patterns of molecular be- 
havior are known and studied; the other applies to a situ- 
ation largely defined by our ignorance of these patterns. 
But the analogy to atomic complementarity is only par- 
tial, because there is nothing in the classical dynamics 
which underlies kinetic theory to suggest that the behavior 
of a gas would be any different if we had performed the 


78 SCIENCE AND THE COMMON UNDERSTANDING 


immense job of locating and measuring what all the mole- 
cules were doing. We might then, it is true, not find it 
natural to talk about temperature, because we would need 
no average behavior; we would have an actual one; but 
we could still define the temperature in terms of the total 
kinetic energy of the molecules, and we would still find 
that it tended to equalize between one part of the system 
and another. 

We have therefore a situation in which there are two 
ways of describing a system, two sets of concepts, two 
centers of preoccupation. One is appropriate when we are 
dealing with a very few molecules and want to know what 
those molecules do; the other appropriate when we have a 
large mass of matter and only rough and large-scale ob- 
servations about it. 

There is, however, no logical or inherent difficulty 
within the framework of classical physics, in combining 
both descriptions for a single system—and classical phys- 
ics, we repeat, is adequate for most, if not all, of these 
problems of statistical mechanics. It is not that we cannot 
do this without violating the laws of physics; it is that it 
makes no sense to do it, since each description is appropri- 
ate to a context quite different from the other. It is clear 
that, if we insisted on the detailed description of the mo- 
tion of individual molecules, the notions of probability 
which turn out to be so essential for our understanding of 
the irreversible character of physical events in nature 
would never enter. We should not have the great insight 
that we now do: namely, that the direction of change in the 
world is from the less probable to the more, from the more 


UNCOMMON SENSE 79 


organized to the less, because all we would be talking 
about would be an incredible number of orbits and trajec- 
tories and collisions. It would be a great miracle to us that, 
out of equations of motion, which to every allowed motion 
permit a precisely opposite one, we could nevertheless 
emerge into a world in which there is a trend of change 
with time which is irreversible, unmistakable, and fa- 


miliar in all our physical experience. 


In considering the relations between the various 
sciences, there are similar instances of complementary 
views. In many cases, it is not clear whether this is the sort 
of complementarity that we have between the statistical 
and dynamic descriptions of a gas, a contrast of interest 
and terminology, but not an inherent inapplicability of 
two ways of talking; or whether on the contrary the situa- 
tion is in fact more as it is in atomic physics, where the 
nature of the world is such that the two modes of descrip- 
tion cannot be applied at once to the same situation. Every 
science has its own language. But dictionaries of transla- 
tion between the languages do exist, and mark an ever- 
growing understanding and unity of science as a whole. 
It is not always clear whether the dictionaries will be 
complete; between physics and chemistry they apparently 
are. Everything the chemist observes and describes can be 
talked about in terms of atomic mechanics, and most of it 
at least can be understood. Yet no one suggests that, in 
dealing with the complex chemical forms which are of 
biological interest, the language of atomic physics would 
be helpful. Rather it would tend to obscure the great 


80 SCIENCE AND THE COMMON UNDERSTANDING 


regularities of biochemistry, as the dynamic description 
of a gas would obscure its thermodynamic behavior. 

The contrast becomes even more marked when we con- 
sider the physico-chemical description of living forms. 
Here, in spite of the miraculous sharpness of the tools of 
chemical analysis, of the extensive use not only of the 
microscope but of the electron microscope to determine 
fine details of biological structure, in spite of the use of 
tracers to follow changes on a molecular scale, questions 
have still been raised as to whether this description can 
in the nature of things be complete. 

The question involves two points: the first having to do 
with the impossibility of wholly isolating a biological sys- 
tem from its physical environment without killing it; the 
second with the possibility that a really complete physico- 
chemical study of the pivotal structures in biological proc- 
esses—of genes, let us say, in the nuclei of dividing cells 
—might not be incompatible with the undisturbed course 
of life itself. It would appear to be the general opinion of 
biologists that no such limitations will prove decisive; that 
a complete description of biology will be possible not only 
in terms of the concepts of biology but in terms reducible 
to those of physics and chemistry. Certainly it is a large 
part of the aim and wonder of biological progress to carry 
this program as far as possible. 

Analogous questions appear much sharper, and their 
answer more uncertain, when we think of the phenomena 
of consciousness; and, despite all the progress that has 
been made in the physiology of the sense organs and of 
the brain, despite our increasing knowledge of these in- 


UNCOMMON SENSE 81 


tricate marvels both as to their structure and their func- 
tioning, it seems rather unlikely that we shall be able to 
describe in physico-chemical terms the physiological 
phenomena which accompany a conscious thought, or 
sentiment, or will. Today the outcome is uncertain. What- 
ever the outcome, we know that, should an understanding 
of the physical correlate of elements of consciousness in- 
deed be available, it will not itself be the appropriate 
description for the thinking man himself, for the clarifica- 
tion of his thoughts, the resolution of his will, or the de- 
light of his eye and mind at works of beauty. Indeed, an 
understanding of the complementary nature of conscious 
life and its physical interpretation appears to me a lasting 
element in human understanding and a proper formula- 
tion of the historic views called psycho-physical paral- 
lelism. 

For within conscious life, and in its relations with the 
description of the physical world, there are again many 
examples. There is the relation between the cognitive and 
the affective sides of our lives, between knowledge or 
analysis and emotion or feeling. There is the relation be- 
tween the aesthetic and the heroic, between feeling and 
that precursor and definer of action, the ethical commit- 
ment; there is the classical relation between the analysis 
of one’s self, the determination of one’s motives and pur- 
poses, and that freedom of choice, that freedom of deci- 
sion and action, which are complementary to it. 

Whether a physico-chemical description of the material 
counterpart of consciousness will in fact ever be possible, 
whether physiological or psychological observation will 


82 SCIENCE AND THE COMMON UNDERSTANDING 


ever permit with any relevant confidence the prediction of 
our behavior in moments of decision and in moments of 
challenge, we may be sure that these analyses and these 
understandings, even should they exist, will be as irrele- 
vant to the acts of decision and the castings of the will as 
are the trajectories of molecules to the entropy of a gas. 
To be touched with awe, or humor, to be moved by beauty, 
to make a commitment or a determination, to understand 
some truth—these are complementary modes of the human 
spirit. All of them are part of man’s spiritual life. None 
can replace the others, and where one is called for the 
others are in abeyance. 

Just as with the a-particles of Rutherford, which were 
first for him an object of study and then became for him 
a tool of study, a tool for investigating other objects, so 
our thoughts and words can be the subject of reflection and 
analysis; so we can be introspective, critical, and full of 
doubt. And so, in other times and other contexts, these 
same words, these same thoughts taken as instruments, are 
the power of human understanding itself, and the means of 
our further enlightenment. 

The wealth and variety of physics itself, the greater 
wealth and variety of the natural sciences taken as a whole, 
the more familiar, yet still strange and far wider wealth 
of the life of the human spirit, enriched by complemen- 
tary, not at once compatible ways, irreducible one to the 
other, have a greater harmony. They are the elements of 
man’s sorrow and his splendor, his frailty and his power, 
his death, his passing, and his undying deeds. 


THE SCIENCES 


AND MAN’S COMMUNITY 


FOR some moments during these lectures we have 
looked together into one of the rooms of the house called 
“science.”” This is a relatively quiet room that we know as 
quantum theory or atomic theory. The great girders which 
frame it, the lights and shadows and vast windows—these 
were the work of a generation our predecessor more than 
two decades ago. It is not wholly quiet. Young people visit 
it and study in it and pass on to other chambers; and from 
time to time someone rearranges a piece of the furniture 
to make the whole more harmonious; and many, as we 
have done, peer through its windows or walk through it 
as sight-seers. It is not so old but that one can hear the 
sound of the new wings being built nearby, where men 
walk high in the air to erect new scaffoldings, not uncon- 
scious of how far they may fall. All about there are busy 
workshops where the builders are active, and very near 
indeed are those of us who, learning more of the primor- 


84 SCIENCE AND THE COMMON UNDERSTANDING 


dial structure of matter, hope some day for chambers as 
fair and lovely as that in which we have spent the years of 
our youth and our prime. 

It is a vast house indeed. It does not appear to have 
been built upon any plan but to have grown as a great city 
grows. There is no central chamber, no one corridor from 
which all others debouch. All about the periphery men are 
at work studying the vast reaches of space and the state of 
affairs billions of years ago; studying the intricate and 
subtle but wonderfully meet mechanisms by which life 
proliferates, alters, and endures; studying the reach of the 
mind and its ways of learning; digging deep into the atoms 
and the atoms within atoms and their unfathomed order. 
It is a house so vast that none of us know it, and even the 
most fortunate have seen most rooms only from the out- 
side or by a fleeting passage, as in a king’s palace open 
to visitors. It is a house so vast that there is not and need 
not be complete concurrence on where its chambers stop 
and those of the neighboring mansions begin. 

It is not arranged in a line nor a square nor a circle nor 
a pyramid, but with a wonderful randomness suggestive 
of unending growth and improvisation. Not many people 
live in the house, relatively speaking—perhaps if we count 
all its chambers and take residence requirements quite 
lightly, one tenth of one per cent, of all the people in this 
world—probably, by any reasonable definition, far fewer. 
And even those who live here live elsewhere also, live in 
houses where the rooms are not labelled atomic theory or 
genetics or the internal constitution of the stars, but quite 


THE SCIENCES AND MAN’S COMMUNITY 85 


different names like power and production and evil and 
beauty and history and children and the word of God. 

We go in and out; even the most assiduous of us is not 
bound to this vast structure. One thing we find throughout 
the house: there are no locks; there are no shut doors; 
wherever we go there are the signs and usually the words 
of welcome. It is an open house, open to all comers. 

The discoveries of science, the new rooms in this great 
house, have changed the way men think of things outside 
its walls. We have some glimmering now of the depth in 
time and the vastness in space of the physical world we 
live in. An awareness of how long our history and how 
immense our cosmos touches us even in simple earthly 
deliberations. We have learned from the natural history of 
the earth and from the story of evolution to have a sense of 
history, of time and change. We learn to talk of ourselves, 
and of the nature of the world and its reality as not wholly 
fixed in a silent quiet moment, but as unfolding with nov- 
elty and alteration, decay and new growth. We have under- 
stood something of the inner harmony and beauty of 
strange primitive cultures, and through this see the qualli- 
ties of our own life in an altered perspective, and recog- 
nize its accidents as well as its inherent necessities. We 
are, I should think, not patriots less but patriots very 
differently for loving what is ours and understanding a 
little of the love of others for their lands and ways. We 
have begun to understand that it is not only in his rational 
life that man’s psyche is intelligible, that even in what may 
appear to be his least rational actions and sentiments we 


86 SCIENCE AND THE COMMON UNDERSTANDING 


may discover a new order. We have the beginnings of an 
understanding of what it is in man, and more in simple 
organisms, that is truly heritable, and rudimentary clues 
as to how the inheritance occurs. We know, in surprising 
detail, what is the physical counterpart of the act of vision 
and of other modes of perception. Not one of these new 
ideas and new insights is so little, or has so short a reach 
in its bearing on the common understanding but that it 
alone could make a proper theme for “Science and the 
Common Understanding.” Yet we have been, bearing in 
mind my limited area of experience, in that one room of 
the part of the house where physics is, in which I have 
for some years worked and taught. 

In that one room—in that relatively quiet room where 
we have been together—we have found things quite 
strange for those who have not been there before, yet 
reminiscent of what we have seen in other houses and 
known in other days. We have seen that in the atomic 
world we have been led by experience to use descriptions 
and ideas that apply to the large-scale world of matter, to 
the familiar world of our schoolday physics; ideas like 
the position of a body and its acceleration and its impulse 
and the forces acting on it; ideas like wave and interfer- 
ence; ideas like cause and probability. But what is new, 
what was not anticipated a half-century ago, is that, though 
to an atomic system there is a potential applicability of 
one or another of these ideas, in any real situation only 
some of these ways of description can be actual. This is 
because we need to take into account not merely the atomic 


THE SCIENCES AND MAN’S COMMUNITY 87 


system we are studying, but the means we use in observing 
it, and the fitness of these experimental means for defining 
and measuring selected properties of the system. All such 
ways of observing are needed for the whole experience of 
the atomic world; all but one are excluded in any actual 
experience. In the specific instance, there is a proper and 
consistent way to describe what the experience is; what it 
implies; what it predicts and thus how to deal with its 
consequences. But any such specific instance excludes by 
its existence the application of other ideas, other modes of 
prediction, other consequences. They are, we say, com- 
plementary to one another; atomic theory is in part an 
account of these descriptions and in part an understanding 
of the circumstances to which one applies, or another or 
another. 

And so it is with man’s life. He may be any of a number 
of things; he will not be all of them. He may be well 
versed, he may be a poet, he may be a creator in one or 
more than one science; he will not be all kinds of man or 
all kinds of scientist; and he will be lucky if he has a bit 
of familiarity outside the room in which he works. 

So it is with the great antinomies that through the ages 
have organized and yet disunited man’s experience: the 
antinomy between the ceaseless change and wonderful 
novelty and the perishing of all earthly things, and the 
eternity which inheres in every happening; in the an- 
tinomy between growth and order, between the spontane- 
ous and changing and irregular and the symmetrical and 
balanced; in the related antinomy between freedom and 


88 SCIENCE AND THE COMMON UNDERSTANDING 


necessity ; between action, the life of the will, and observa- 
tion and analysis and the life of reason; between the ques- 
tion “how?” and the questions “why?” and “to what 
end?’’; between the causes that derive from natural law, 
from unvarying regularities in the natural world, and 
those other causes that express purposes and define goals 
and ends. 


So it is in the antinomy between the individual and the 
community; man who is an end in himself and man whose 
tradition, whose culture, whose works, whose words have 
meaning in terms of other men and his relations to them. 
All our experience has shown that we can neither think, 
nor in any true sense live, without reference to these anti- 
nomic modes. We cannot in any sense be both the observ- 
ers and the actors in any specific instance, or we shall fail 
properly to be either one or the other; yet we know that 
our life is built of these two modes, is part free and part 
inevitable, is part creation and part discipline, is part ac- 
ceptance and part effort. We have no written rules that 
assign us to these ways; but we know that only folly and 
death of the spirit results when we deny one or the other, 
when we erect one as total and absolute and make the 
others derivative and secondary. We recognize this when 
we live as men. We talk to one another; we philosophize; 
we admire great men and their moments of greatness; we 
read; we study; we recognize and love in a particular act 
that happy union of the generally incompatible. With all 


THE SCIENCES AND MAN’S COMMUNITY 89 


of this we learn to use some reasonable part of the full 
register of man’s resources. 

We are, of course, an ignorant lot; even the best of us 
knows how to do only a very few things well; and of what 
is available in knowledge of fact, whether of science or of 
history, only the smallest part is in any one man’s know- 
ing. 

The greatest of the changes that science has brought is 
the acuity of change; the greatest novelty the extent of 
novelty. Short of rare times of great disaster, civilizations 
have not known such rapid alteration in the conditions of 
their life, such rapid flowering of many varied sciences, 
such rapid changes in the ideas we have about the world 
and one another. What has been true in the days of a great 
disaster or great military defeat for one people at one 
time is true for all of us now, in the sense that our ends 
have little in common with our beginnings. Within a life- 
time what we learned at school has been rendered inade- 
quate by new discoveries and new inventions; the ways 
that we learn in childhood are only very meagerly ade- 
quate to the issues that we must meet in maturity. 

In fact, of course, the notion of universal knowledge has 
always been an illusion; but it is an illusion fostered by 
the monistic view of the world in which a few great central 
truths determine in all its wonderful and amazing prolif- 
eration everything else that is true. We are not today 
tempted to search for these keys that unlock the whole of 
human knowledge and of man’s experience. We know that 
we are ignorant; we are well taught it, and the more 


go SCIENCE AND THE COMMON UNDERSTANDING 


surely and deeply we know our own job the better able 
we are to appreciate the full measure of our pervasive 
ignorance. We know that these are inherent limits, com- 
pounded, no doubt, and exaggerated by that sloth and that 
complacency without which we would not be men at all. 

But knowledge rests on knowledge; what is new is mean- 
ingful because it departs slightly from what was known 
before; this is a world of frontiers, where even the live- 
liest of actors or observers will be absent most of the time 
from most of them. Perhaps this sense was not so sharp in 
the village—that village which we have learned a little 
about but probably do not understand too well—the vil- 
lage of slow change and isolation and fixed culture which 
evokes our nostalgia even if not our full comprehension. 
Perhaps in the villages men were not so lonely; perhaps 
they found in each other a fixed community, a fixed and 
only slowly growing store of knowledge—a single world. 
Even that we may doubt, for there seem to be always in the 
culture of such times and places vast domains of mystery, 
if not unknowable, then imperfectly known, endless and 
open. 


As for ourselves in these times of change, of ever-in- 
creasing knowledge, of collective power and individual 
impotence, of heroism and of drudgery, of progress and of 
tragedy, wetoo are brothers. And if we, who are the inher- 
itors of two millennia of Christian tradition, understand 


that for us we have come to be brothers second by being 


THE SCIENCES AND MAN’S COMMUNITY g1 


children first, we know that in vast parts of the world 
where there has been no Christian tradition, and with men 
who never have been and never may be Christian in faith 
there is nevertheless a bond of brotherhood. We know 
this not only because of the almost universal ideal of 
human brotherhood and human community; we know it at 
first hand from the more modest, more diverse, more fleet- 
ing associations which are the substance of our life. The 
ideal of brotherhood, the ideal of fraternity in which all 
men, wicked and virtuous, wretched and fortunate, are 
banded together has its counterpart in the experience of 
communities, not ideal, not universal, imperfect, imper- 
manent, as different from the ideal and as reminiscent of 
it as are the ramified branches of science from the ideal of 
a unitary, all-encompassing science of the eighteenth cen- 
tury. 

Each of us knows from his own life how much even a 
casual and limited association of men goes beyond him in 
knowledge, in understanding, in humanity, and in power. 
Each of us, from a friend or a book or by concerting of 
the little we know with what others know, has broken the 
iron circle of his frustration. Each of us has asked help 
and been given it, and within our measure each of us has 
offered it. Each of us knows the great new freedom sensed 
almost as a miracle, that men banded together for some 
finite purpose experience from the power of their common 
effort. We are likely to remember the times of the last war, 
where the common danger brought forth in soldier, in 
worker, in scientist, and engineer a host of new experi- 


g2 SCIENCE AND THE COMMON UNDERSTANDING 


ences of the power and the comfort in even bleak under- 
takings, of common, concerted, co-operative life. Each of 
us knows how much he has been transcended by the group 
of which he has been or is a part; each of us has felt the 
solace of other men’s knowledge to stay his own ignorance, 
of other men’s wisdom to stay his folly, of other men’s 
courage to answer his doubts or his weakness. 

These are the fluid communities, some of long duration 
when circumstances favored—like the political party or 
many a trade union—some fleeting and vivid, encompass- 
ing in the time of their duration a moment only of the 
member’s life; and in our world at least they are ramified 
and improvised, living and dying, growing and falling off 
almost as a form of life itself. This may be more true of 
the United States than of any other country. Certainly the 
bizarre and comical aspects impressed de Tocqueville 
more than a century ago when he visited our land and com- 
mented on the readiness with which men would band to- 
gether: to improve the planting of a town, or for political 
reform, or for the pursuit or inter-exchange of knowledge, 
or just for the sake of banding together, because they 
liked one another or disliked someone else. Circumstances 
may have exaggerated the role of the societies, of the fluid 
and yet intense communities in the United States; yet these 
form a common pattern for our civilization. It brought 
men together in the Royal Society and in the French 
Academy and in the Philosophical Society that Franklin 
founded, in family, in platoon, ona ship, in the laboratory, 
in almost everything but a really proper club. 


THE SCIENCES AND MAN’S COMMUNITY 93 


If we err today—and I think we do—it is in expecting 
too much of knowledge from the individual and too much 
of synthesis from the community. We tend to think of these 
communities, no less than of the larger brotherhood of 
man, as made up of individuals, as composed of them as 
an atom is of its ingredients. We think similarly of general 
laws and broad ideas.as made up of the instances which 
illustrate them, and from an observation of which we may 
have learned them. 

Yet this is not the whole. The individual event, the act, 
goes far beyond the general law. It is a sort of intersection 
of many generalities, harmonizing them in one instance as 
they cannot be harmonized in general. And we as men are 
not only the ingredients of our communities; we are their 
intersection, making a harmony which does not exist be- 
tween the communities except as we, the individual men, 
may create it and reveal it. So much of what we think, our 
acts, our judgments of beauty and of right and wrong, 
come to us from our fellow men that what would be left 
were we to take all this away would be neither recogniza- 
ble nor human. We are men because we are part of, but 
not because only part of, communities; and the attempt to 
understand man’s brotherhood in terms only of the indi- 
vidual man is as little likely to describe our world as is the 
attempt to describe general laws as the summary of their 
instances. These are indeed two complementary views, 
neither reducible to the other, no more reducible than is 
the electron as wave to the electron as particle. 

And this is the mitigant of our ignorance. It is true that 


94 SCIENCE AND THE COMMON UNDERSTANDING 


none of us will know very much; and most of us will see 
the end of our days without understanding in all its detail 
and beauty the wonders uncovered even in a single branch 
of a single science. Most of us will not even know, as a 
member of any intimate circle, anyone who has such 
knowledge; but it is also true that, although we are sure 
not to know everything and rather likely not to know very 
much, we can know anything that is known to man, and 
may, with luck and sweat, even find out some things that 
have not before been known to him. This possibility, 
which, as a universal condition of man’s life is new, repre- 
sents today a high and determined hope, not yet a reality; 
it is for us in England and in the United States not wholly 
remote or unfamiliar. It is one of the manifestations of our 
belief in equality, that belief which could perhaps better 
be described as a commitment to unparalleled diversity 
and unevenness in the distribution of attainments, knowl- 
edge, talent, and power. 

This open access to knowledge, these unlocked doors 
and signs of welcome, are a mark of a freedom as funda- 
mental as any. They give a freedom to resolve difference 
by converse, and, where converse does not unite, to let 
tolerance compose diversity. This would appear to be a 
freedom barely compatible with modern political tyranny. 
The multitude of communities, the free association for 
converse or for common purpose, are acts of creation. It 
is not merely that without them the individual is the 
poorer; without them a part of human life, not more nor 
less fundamental than the individual, is foreclosed. It is a 


THE SCIENCES AND MAN’S COMMUNITY 95 


cruel and humorless sort of pun that so powerful a present 
form of modern tyranny should call itself by the very 
name of a belief in community, by a word “communism” 
which in other times evoked memories of villages and vil- 
lage inns and of artisans concerting their skills, and of 
men of learning content with anonymity. But perhaps 
only a malignant end can follow the systematic belief that 
all communities are one community; that all truth is one 
truth; that all experience is compatible with all other; that 
total knowledge is possible; that all that is potential can 
exist as actual. This is not man’s fate; this is not his path; 
to force him on it makes him resemble not that divine 
image of the all-knowing and all-powerful but the help- 
less, iron-bound prisoner of a dying world. The open 
society, the unrestricted access to knowledge, the un- 
planned and uninhibited association of men for its further- 
ance—these are what may make a vast, complex, ever- 
growing, ever-changing, ever more specialized and expert 
technological world nevertheless a world of human com- 


munity. 


So it is with the unity of science—that unity that is far 
more a unity of comparable dedication than a unity of 
common total understanding. This heartening phrase, “the 
unity of science,” often tends to evoke a wholly false pic- 
ture, a picture of a few basic truths, a few critical tech- 
niques, methods, and ideas, from which all discoveries 


and understanding of science derive; a sort of central ex- 


96 SCIENCE AND THE COMMON UNDERSTANDING 


change, access to which will illuminate the atoms and the 
galaxies, the genes and the sense organs. The unity of 
science is based rather on just such a community as I have 
described. All parts of it are open to all of us, and this is 
no merely formal invitation. The history of science is rich 
in example of the fruitfulness of bringing two sets of tech- 
niques, two sets of ideas, developed in separate contexts 
for the pursuit of new truth, into touch with one another. 
The sciences fertilize each other; they grow by contact and 
by common enterprise. Once again, this means that the 
scientist may profit from learning about any other science; 
it does not mean that he must learn about them all. It 
means that the unity is a potential unity, the unity of the 
things that might be brought together and might throw 
light one on the other. It is not global or total or hier- 
archical. 

Even in science, and even without visiting the room in 
its house called atomic theory, we are again and again 
reminded of the complementary traits in our own life, 
even in our own professional life. We are nothing without 
the work of others our predecessors, others our teachers, 
others our contemporaries. Even when, in the measure of 
our adequacy and our fullness, new insight and new order 
are created, we are still nothing without others. Yet we 
are more. 

There is a similar duality in our relations to wider 
society. For society our work means many things: pleas- 
ure, we hope, for those who follow it; instruction for those 
who perhaps need it; but also and far more widely, it 


THE SCIENCES AND MAN’S COMMUNITY 97 


means a common power, a power to achieve that which 
could not be achieved without knowledge. It means the 
cure of illness and the alleviation of suffering; it means 
the easing of labor and the widening of the readily acces- 
sible frontiers of experience, of communication, and of 
instruction. It means, in an earthy way, the power of better- 
ment—that riddled word. We are today anxiously aware 
that the power to change is not always necessarily good. 

As new instruments of war, of newly massive terror, 
add to the ferocity and totality of warfare, we understand 
that it is a special mark and problem of our age that man’s 
ever-present preoccupation with improving his lot, with 
alleviating hunger and poverty and exploitation, must be 
brought into harmony with the over-riding need to limit 
and largely to eliminate resort to organized violence be- 
tween nation and nation. The increasingly expert destruc- 
tion of man’s spirit by the power of police, more wicked if 
not more awful than the ravages of nature’s own hand, is 
another such power, good only if never to be used. 

We regard it as proper and just that the patronage of 
science by society is in large measure based on the in- 
creased power which knowledge gives. If we are anxious 
that the power so given and so obtained be used with wis- 
dom and with love of humanity, that is an anxiety we 
share with almost everyone. But we also know how little 
of the deep new knowledge which has altered the face of 
the world, which has changed—and increasingly and ever 
more profoundly must change—man’s views of the world, 
resulted from a quest for practical ends or an interest in 


98 SCIENCE AND THE COMMON UNDERSTANDING 


exercising the power that knowledge gives. For most of 
us, in most of those moments when we were most free of 
corruption, it has been the beauty of the world of nature 
and the strange and compelling harmony of its order, that 
has sustained, inspirited, and led us. That also is as it 
should be. And if the forms in which society provides and 
exercises its patronage leave these incentives strong and 
secure, new knowledge will never stop as long as there are 
men. 

We know that our work is rightly both an instrument 
and an end. A great discovery is a thing of beauty; and 
our faith—our binding, quiet faith—is that knowledge is 
good and good in itself. It is also an instrument; it is an 
instrument for our successors, who will use it to probe else- 
where and more deeply; it is an instrument for technology, 
for the practical arts, and for man’s affairs. So it is with 
us as scientists; so it is with us as men. We are at once 
instrument and end, discoverers and teachers, actors and 
observers. We understand, as we hope others understand, 
that in this there is a harmony between knowledge in the 
sense of science, that specialized and general knowledge 
which it is our purpose to uncover, and the community of 
man. We, like all men, are among those who bring a little 
light to the vast unending darkness of man’s life and 
world. For us as for all men, change and eternity, special- 
ization and unity, instrument and final purpose, com- 
munity and individual man alone, complementary each 
to the other, both require and define our bonds and our 
freedom. 


NOTE 


The six chapters of this book are the Reith Lectures given 
over the home service of the British Broadcasting Corpora- 
tion in November and December, 1953. They are printed es- 
sentially as broadcast. I have added two appendixes. 

In the one I have collected the texts from which brief 
quotation is made in the lectures. The texts seem to me of 
interest in themselves; and in any case, this is the surest way 
to correct any distortion or color which my abbreviation may 
have introduced. In some cases the texts may even prompt 
curiosity to read further. 

In the second appendix, I have given a brief and informal 
bibliography on atomic theory and its interpretation. 


J. R. O. 


APPENDIX I 


SIR ISAAC NEWTON 
(Page 12) 


All these things being consider’d, it seems probable to me, 
that God in the Beginning form’d Matter in solid, massy, 
hard, impenetrable, moveable Particles, of such Sizes and 
Figures, and with such other Properties, and in such Propor- 
tion to Space, as most conduced to the End for which he 
form’d them; and that these primitive Particles being Solids, 
are incomparably harder than any porous Bodies compounded 
of them; even so very hard, as never to wear or break in 
pieces; no ordinary Power being able to divide what God him- 
self made one in the first Creation. While the Particles con- 
tinue entire, they may compose Bodies of one and the same Na- 
ture and Texture in all Ages: But should they wear away, or 
break in pieces, the Nature of Things depending on them, 
would be changed. Water and Earth, composed of old worn 
Particles and Fragments of Particles, would not be of the 
same Nature and Texture now, with Water and Earth com- 
posed of entire Particles in the Beginning. And therefore, 
that Nature may be lasting, the Changes of corporeal Things 


SCIENCE AND THE COMMON UNDERSTANDING 103 


are to be placed only in the various Separations and new 
Associations and Motions of these permanent Particles; com- 
pound Bodies being apt to break, not in the midst of solid 
Particles, but where those Particles are laid together, and 
only touch in a few Points. 

It seems to me farther, that these Particles have not only a 
Vis inerte, accompanied with such passive Laws of Motion 
as naturally result from that Force, but also that they are 
moved by certain active Principles, such as is that of Gravity, 
and that which causes Fermentation, and the Cohesion of 
Bodies. These Principles I consider, not as occult Qualities, 
supposed to result from the specifick Forms of Things, but 
as general Laws of Nature, by which the Things themselves 
are form’d; their Truth appearing to us by Phenomena, 
though their Causes be not yet discover’d. . . . 


Sir Isaac Newton, Opticks (New York: Dover 
Publications, Inc., 1952), Book 3, Part I, Query 
31, p. 400. Based on the Fourth Edition, London, 
1730. 


THOMAS SPRAT 
(Page 16) 


I will here, in the first place, contract into few Words, the 
whole Sum of their Resolutions; which I shall often have 
occasion to touch upon in Parcels. Their Purpose is, in short, 
to make faithful Records of all the Works of Nature, or Art, 
which can come within their Reach; that so the present Age, 
and Posterity, may be able to put a Mark on the Errors, 
which have been strengthned by long Prescription; to restore 
the Truths, that have lain neglected; to push on those, which 
are already known, to more various Uses; and to make the 
way more passable, to what remains unreveal’d. This is the 
Compass of their Design. And to accomplish this, they have 
endeavour’d, to separate the Knowledge of Nature, from the 
Colours Rhetorick, the Devices of Fancy, or the delightful 
Deceit of Fables. They have labor’d to inlarge it, from being 
confin’d to the Custody of a few, or from Servitude to private 
Interests. They have striven to preserve it from being over- 
press’d by a confus’d Heap of vain and useless Particulars; 
or from being streightned and bound too much up by general 


SCIENCE AND THE COMMON UNDERSTANDING 105 


Doctrines. They have tried to put it into a Condition of per- 
petual Increasing; by settling an inviolable Correspondence 
between the Hand and the Brain. They have studied, to make 
it not only an Enterprise of one Season, or of some lucky 
Opportunity; but a Business of Time; a steady, a lasting, a 
popular, an uninterrupted Work. They have attempted, to 
free it from the Artifice, and Humors, and Passions of Sects; 
to render it an Instrument, whereby Mankind may obtain a 
Dominion over Things, and not only over one another’s Judg- 
ments: And lastly, they have begun to establish these Refor- 
mations in Philosophy, not so much, by any solemnity of 
Laws, or Ostentation of Ceremonies, as by solid Practice and 
Examples; not by a glorious Pomp of Words; but by the si- 
lent, effectual, and unanswerable Arguments of real Produc- 
tions. 

This will more fully appear, by what I am to say on these 
four Particulars, which shall make up this Part of my Rela- 
tion, the Qualifications of their Members; the Manner of their 
Inquiry; their Weekly Assemblies; and their Way of Reg- 
istring. 

As for what belongs to the Members themselves that are to 
constitute the Society: It is to be noted, that they have freely 
admitted Men of different Religions, Countries, and Profes- 
sions of Life. This they were oblig’d to do, or else they would 
come far short of the Largeness of their own Declarations. 
For they openly profess, not to lay the Foundation of an Eng- 
lish, Scotch, Irish, Popish, or Protestant Philosophy; but a 
Philosophy of Mankind. 

That the Church of England ought not to be apprehensive 
of this free Converse of various Judgments, I shall afterwards 
manifest at large. For the present, I shall frankly assert, that 


106 APPENDIX I 


our Doctrine, and Discipline, will be so far from receiving 
Damage by it; that it were the best Way to make them uni- 
versally embrac’d, if they were oftner brought to be canvass’d 
amidst all Sorts of Dissenters. It is dishonorable, to pass a 
hard Censure on the Religions of all other Countries: It con- 
cerns them, to look to the Reasonableness of their Faith; and 
it is sufficient for us, to be establish’d in the Truth of our own. 


Thomas Sprat, The History of the Royal Society of 
London (3rd ed.; London, 1722), pp. 61-63. 


THOMAS JEFFERSON 
(Page 18) 


Monticello June 18. 99. 
DEAR SIR, 

I have to acknolege the reciept of your favor of May 14. 
in which you mention that you have finished the 6. first books 
of Euclid, plane trigonometry, surveying and algebra and 
ask whether I think a further pursuit of that branch of science 
would be useful to you. There are some propositions in the 
latter books of Euclid, and some of Archimedes, which are 
useful, and I have no doubt you have been made acquainted 
with them. Trigonometry, so far as this, is most valuable to 
every man, there is scarcely a day in which he will not resort 
to it for some of the purposes of common life; the science of 
calculation also is indispensible as far as the extraction of 
the square and cube roots; Algebra as far as the quadratic 
equation and the use of logarithms are often of value in ordi- 
nary cases: but all beyond these is but a luxury; a delicious 
luxury indeed; but not to be indulged in by one who is to 
have a profession to follow for his subsistence. In this light 


108 APPENDIX I 


I view the conic sections, curves of the higher orders, perhaps 
even spherical trigonometry, Algebraical operations beyond 
the 2d dimension, and fluxions. There are other branches of 
science however worth the attention of every man: Astron- 
omy, botany, chemistry, natural philosophy, natural history, 
anatomy. Not indeed to be a proficient in them; but to pos- 
sess their general principles and outlines, so as that we may 
be able to amuse and inform ourselves further in any of them 
as we proceed through life and have occasion for them. Some 
knowlege of them is necessary for our character as well as 
comfort. The general elements of astronomy and of natural 
philosophy are best acquired at an academy where we can 
have the benefit of the instruments and apparatus usually 
provided there: but the others may well be acquired from 
books alone as far as our purposes require. I have indulged 
myself in these observations to you, because the evidence 
cannot be unuseful to you of a person who has often had 
occasion to consider which of his acquisitions in science have 
been really useful to him in life, and which of them have been 
merely a matter of luxury. 

I am among those who think well of the human character 
generally. I consider man as formed for society, and endowed 
by nature with those dispositions which fit him for society. 
I believe also, with Condorcet, as mentioned in your letter, 
that his mind is perfectible to a degree of which we cannot 
as yet form any conception. It is impossible for a man who 
takes a survey of what is already known, not to see what an 
immensity in every branch of science yet remains to be dis- 
covered, and that too of articles to which our faculties seem 
adequate. In geometry and calculation we know a great deal. 
Yet there are some desiderata. In anatomy great progress has 


SCIENCE AND THE COMMON UNDERSTANDING 109 


been made; but much is still to be acquired. In natural his- 
tory we possess knowlege; but we want a great deal. In chem- 
istry we are not yet sure of the first elements. Our natural 
philosophy is in a very infantine state; perhaps for great ad- 
vances in it, a further progress in chemistry is necessary. 
Surgery is well advanced; but prodigiously short of what may 
be. The state of medecine is worse than that of total ignor- 
ance. Could we divest ourselves of every thing we suppose 
we know in it, we should start from a higher ground and with 
fairer prospects. From Hippocrates to Brown we have had 
nothing but a succession of hypothetical systems each having 
it’s day of vogue, like the fashions and fancies of caps and 
gowns, and yielding in turn to the next caprice. Yet the hu- 
man frame, which is to be the subject of suffering and torture 
under these learned modes, does not change. We have a few 
medecines, as the bark, opium, mercury, which in a few well 
defined diseases are of unquestionable virtue: but the resid- 
uary list of the materia medica, long as it is, contains but the 
charlataneries of the art; and of the diseases of doubtful form, 
physicians have ever had a false knowlege, worse than ignor- 
ance. Yet surely the list of unequivocal diseases and remedies 
is capable of enlargement; and it is still more certain that in 
the other branches of science, great fields are yet to be ex- 
plored to which our faculties are equal, and that to an extent 
of which we cannot fix the limits. I join you therefore in 
branding as cowardly the idea that the human mind is in- 
capable of further advances. This is precisely the doctrine 
which the present despots of the earth are inculcating, and 
their friends here re-echoing; and applying especially to re- 
ligion and politics; ‘that it is not probable that any thing 
better will be discovered than what was known to our fathers.” 


110 APPENDIX I 


We are to look backwards then and not forwards for the im- 
provement of science, and to find it amidst feudal barbarisms 
and the fires of Spital-fields. But thank heaven the American 
mind is already too much opened, to listen to these impos- 
tures; and while the art of printing is left to us, science can 
never be retrograde; what is once acquired of real knowlege 
can never be lost. To preserve the freedom of the human 
mind then and freedom of the press, every spirit should be 
ready to devote itself to martyrdom; for as long as we may 
think as we will, and speak as we think, the condition of man 
will proceed in improvement. The generation which is going 
off the stage has deserved well of mankind for the struggles 
it has made, and for having arrested that course of despotism 
which had overwhelmed the world for thousands and thou- 
sands of years. If there seems to be danger that the ground 
they have gained will be lost again, that danger comes from 
the generation your cotemporary. But that the enthusiasm 
which characterises youth should lift it’s parracide hands 
against freedom and science would be such a monstrous 
phaenomenon as | cannot place among possible things in this 
age and this country. Your college at least has shewn itself 
incapable of it; and if the youth of any other place have 
seemed to rally under other banners it has been from de- 
lusions which they will soon dissipate. I shall be happy to 
hear from you from time to time, and of your progress in 
study, and to be useful to you in whatever is in my power; 
being with sincere esteem Dear Sir 

Your friend & servt 


Th: Jefferson 


Scripta Mathematica, I (1932), 88-90. 


THOMAS HOBBES 
(Page 21) 


Good successe is Power; because it maketh reputation of 
Wisdome, or good fortune; which makes men either feare 
him, or rely on him. 

Affability of men already in power, is encrease of Power; 
because it gaineth love. 

Reputation of Prudence in the conduct of Peace or War, is 
Power; because to prudent men, we commit the government 
of our selves, more willingly than to others. 

Nobility is Power, not in all places, but onely in those 
Common-Wealths, where it has Priviledges: for in such 
priviledges consisteth their Power. 

Eloquence is power; because it is seeming Prudence. 

Forme is Power; because being a promise of Good, it rec- 
ommendeth men to the favour of women and strangers. 

The Sciences, are small Power; because not eminent; and 
therefore, not acknowledged in any man; nor are at all, but in 
a few; and in them, but of a few things. For Science is of that 
nature, as none can understand it to be, but such as in a good 


measure have attayned it. 


112 APPENDIX I 


Arts of publique use, as Fortification, making of Engines, 
and other Instruments of War; because they conferre to De- 
fence, and Victory, are Power: And though the true Mother 
of them, be Science, namely the Mathematiques; yet, because 
they are brought into the Light, by the hand of the Artificer, 
they be esteemed (the Midwife passing with the vulgar for 
the Mother,) as his issue. 


Thomas Hobbes, Leviathan, ed. by A. R. Waller 
(“Cambridge English Classics”; Cambridge: 
Cambridge University Press, 1904), Part I, Chap. 
10, pp. 54-55. 


INSCRIPTION ON STEEPLE 
KNOB OF ST. MARGARET'S 
CHURCH AT GOTHA 


(Page 68) 


Unsere Tage fiillten den gliicklichsten Zeitraum des acht- 
zehnten Jahrhunderts. Kaiser, Kénige, Fiirsten steigen von 
ihrer gefiirchteten Hohe menschenfreundlich herab, vera- 
chten Pracht und Schimmer, werden Vater, Freunde und 
Vertraute ihres Volks. Die Religion zerreisst das Pfaffenge- 
wand und tritt in ihrer Gottlichkeit hervor. Aufklarung geht 
mit Riesenschritten. Tausende unserer Briider und Schwe- 
stern, die in geheiligter Unthatigkeit lebten, werden dem 
Staat geschenkt. Glaubenshass und Gewissenszwang sinken 
dahin; Menschenliebe und Freiheit im Denken gewinnen die 
Oberhand. Kinste und Wissenschaften bliihen, und tief drin- 
gen unsere Blicke in die Werkstatt der Natur. Handwerker 
nahern sich gleich den Kiinstlern der Vollkommenheit, niitz- 
liche Kenntnisse Keimen in allen Standen. Hier habt Ihr eine 
getreue Schilderung unserer Zeit. Blickt nicht stolz auf uns 
herab, wenn Ihr hcher steht und weiter seht als wir; erkennt 
vielmehr aus dem gegebenen Gemilde, wie sehr wir mit Muth 


114 APPENDIX I 


und Kraft Euren Standort emporhoben und stiitzten. Thut 
fir Eure Nachkommenschaft ein Gleiches und seid gliicklich! 


Our days have been the happiest time of the eighteenth 
century. Emperors, kings and princes step down from their 
feared heights, and as friends of men scorn pomp and glitter 
and become fathers, friends and confidants of their people. 
Religion tears off its popish garb and stands forth in its di- 
vinity. Enlightenment advances with giant steps. Thousands 
of our brothers and sisters who previously spent their lives in 
holied idleness are given back to the community. Hatred born 
of dogma and the compulsion of conscience sink away; love 
of man and freedom of thought gain the upper hand. The arts 
and sciences blossom, and our vision into the workshop of 
nature goes deep. Artisans approach artists in perfection; 
useful skills flower at all levels. Here you have a faithful por- 
trait of our time. Look not proudly down upon us, should you 
stand higher or see farther than we, but rather recognize from 
this picture how with courage and strength we raised and 
supported your standard. Do the same for those who come 
after you and rejoice! 


Hermann Hettner, Literaturgeschichte des Acht- 
zehnten Jahrhunderts, Vol. III (Braunschweig: 
Friederich Vieweg und Sohn, 1879), Book 2, 
Chap. 1, p. 171. 


ALEXIS DE TOCQUEVILLE 


(Page 92) 


DE L’USAGE QUE LES AMERICAINS FONT DE 
L’ASSOCIATION DANS LA VIE CIVILE 


Je ne veux point parler de ces associations politiques a 
aide desquelles les hommes cherchent a4 se défendre contre 
Yaction despotique d’une majorité ou contre les empiéte- 
ments du pouvoir royal. J’ai déja traité ce sujet ailleurs. I] 
est clair que si chaque citoyen, a mesure qu’il devient indivi- 
duellement plus faible, et par conséquent plus incapable de 
préserver isolément sa liberté, n’apprenait pas I’art de s’unir 
a ses semblables pour la défendre, la tyrannie croitrait né- 
cessairement avec l’égalité. I] ne s’agit ici que des associa- 
tions qui se forment dans la vie civile, et dont l’objet n’a rien 
de politique. 

Les associations politiques qui existent aux Etats-Unis ne 
forment qu’un détail au milieu de l’immense tableau que 
l’ensemble des associations y présente. 

Les Américains de tous les ages, de toutes les conditions, de 
tous les esprits, s’unissent sans cesse. Non-seulement ils ont 
des associations commerciales et industrielles auxquelles tous 
prennent part, mais ils en ont encore de mille autres espéces: 
de religieuses, de morales, de graves, de futiles, de fort géné- 


116 APPENDIX I 


rales et de trés-particuliéres, d’immenses et de fort petites; les 
Américains s’associent pour donner des fétes, fonder des sé- 
minaires, batir des auberges, élever des églises, répandre des 
livres, envoyer des missionaires aux antipodes; ils créent de 
cette maniére des hdpitaux, des prisons, des écoles. S’agit-il 
enfin de mettre en lumiére une vérité, ou de développer un 
sentiment par l’appui d’un grand exemple: ils s’associent. 
Partout oi, a la téte d’une entreprise nouvelle, vous voyez en 
France le gouvernement, et en Angleterre un grand seigneur, 
comptez que vous apercevrez aux Etats-Unis une association. 
J’ai rencontré en Amérique des sortes d’associations dont 
je confesse que je n’avais pas méme lidée, et j’ai souvent 
admiré l’art infini avec lequel les habitants des Etats-Unis 
parvenaient a fixer un but commun aux efforts d’un grand 
nombre d’hommes, et a les faire marcher librement. 


OF THE USE WHICH THE AMERICANS MAKE OF 
PUBLIC ASSOCIATIONS IN CIVIL LIFE 

I do not propose to speak of those political associations by 
the aid of which men endeavor to defend themselves against 
the despotic action of a majority or against the aggressions of 
regal power. That subject I have already treated. If each 
citizen did not learn, in proportion as he individually be- 
comes more feeble and consequently more incapable of pre- 
serving his freedom singlehanded, to combine with his fel- 
low citizens for the purpose of defending it, it is clear that 
tyranny would unavoidably increase together with equality. 

Only those associations that are formed in civil life without 
reference to political objects are here referred to. The politi- 
cal associations that exist in the United States are only a 
single feature in the midst of the immense assemblage of asso- 


SCIENCE AND THE COMMON UNDERSTANDING 117 


ciations in that country. Americans of all ages, all conditions, 
and all dispositions constantly form associations. They have 
not only commercial and manufacturing companies, in which 
all take part, but associations of a thousand other kinds, re- 
ligious, moral, serious, futile, general or restricted, enormous 
or diminutive. The Americans make associations to give en- 
tertainments, to found seminaries, to build inns, to construct 
churches, to diffuse books, to send missionaries to the an- 
tipodes; in this manner they found hospitals, prisons, and 
schools. If it is proposed to inculcate some truth or to foster 
some feeling by the encouragement of a great example, they 
form a society. Wherever at the head of some new under- 
taking you see the government in France, or a man of rank 
in England, in the United States you will be sure to find an 
association. 

I met with several kinds of associations in America of 
which I confess I had no previous notion; and I have often 
admired the extreme skill with which the inhabitants of the 
United States succeed in proposing a common object for the 
exertions of a great many men and in inducing them volun- 
tarily to pursue it. 


Alexis de Tocqueville, De la Démocratie en Amé- 
rique, Vol. III (14iéme édition; Paris: Michel 
Levy Fréres, 1864), Deuxiéme Partie, Chap. 2, 
p. 175. English translation from Alexis de Tocque- 
ville, Democracy in America, Vol. I] (New York: 
Alfred A. Knopf, 1948), Book 2, Chap. 5, p. 106. 


APPENDIX II 


Chapters Two to Five deal at some length with atomic 
theory and with some of the experiments that underlie it. 
With the exception of contemporary work on the “new parti- 
cles,” there are many admirable technical text books and 
monographs. 

As for Chapter Two, the interested reader may wish to 
turn to the classic texts of E. R. Rutherford, Radioactive 
Substances and their Radiations (Cambridge: Cambridge 
University Press; New York: Putnam; 1913), and to Ruther- 
ford, Chadwick, and Ellis, Radiations from Radioactive Sub- 
stances (Cambridge: Cambridge University Press, 1930). 
The “new particles” were discussed at a conference held at 
Bagnéres de Bigorre, July, 1953. The record of this confer- 
ence, issued by the Ecole Polytechnique, Paris, gives a most 
vivid impression of the present state of knowledge, ignor- 
ance, and progress. Even for those who do not wish to con- 
sult the proceedings of the conference, the comment follow- 
ing the title may be of interest: “Les particules décrites au 
cours de ce Congrés ne sont pas entiérement fictives, et toute 


SCIENCE AND THE COMMON UNDERSTANDING 119 


analogie avec des particules existant dans la nature n’est pas 
une pure coincidence.” 

There are numerous good technical texts on the quantum 
mechanics, the quantum theory of atoms. In particular I 
recommend: 

P. A. M. Dirac. The Principles of Quantum Mechanics. Ox- 

ford: Clarendon Press, 1930. 
W. Pauli. “Die Allgemeinen Prinzipien der Wellenmecha- 
nik,” Handbuch der Physik, XXIV (1933), 1, 83. 
L. I. Schiff. Quantum Mechanics. New York: McGraw-Hill 
Book Co., 1949. 
Of these, Schiff’s text is the most elementary. 

No attempt has been made in the lectures to give a full 
historic account of the contributions made to the develop- 
ment of quantum theory. Where names have been mentioned, 
it is because they have become generally identified with 
principles or with theories; but any account of the history 
of quantum theory should at least mention Born, Dirac, and 
Pauli, in addition to the names that occur in the text. 

As to the interpretation of quantum theory, these may 
serve to guide the reader should he want more detailed, 
more original, and more substantive accounts of the matters 
touched on in the lectures: 

W. Heisenberg. The Physical Principles of the Quantum 

Theory. Chicago: University of Chicago Press, 1930. 
N. Bohr. Atomic Theory and the Description of Nature. 
New York: (Cambridge University Press) The Mac- 
millan Co., 1934. 

N. Bohr. “On the Notions of Causality and Complementar- 
ity,” Dialectica, II (1948), 312. 


120 APPENDIX II 


N. Bohr. “Discussion with Einstein on Epistemological 
Problems in Atomic Physics,” Albert Einstein, Philoso- 
pher-Scientist. Edited by P. A. Schilpp. “Library of 
Living Philosophers”; Evanston, Illinois, 1949. 

W. Pauli. “Die philosophische Bedeutung der Idee der 
Komplementaritat,” Experientia, VI (1950), 72. 


ABOUT THE AUTHOR 


J. RopertT OPPENHEIMER has been director of the Institute 
for Advanced Study at Princeton, New Jersey, since 1947. He 
is a physicist trained at Harvard, Cambridge and Gottingen, 
who has been a professor at the University of California and 
at the California Institute of Technology. Between 1943 and 
1945 he was director of the laboratory at Los Alamos in New 


Mexico, where the first atomic bombs were made.